You are on page 1of 27

STRUCTURAL CONTROL AND HEALTH MONITORING

Struct. Control Health Monit. 2007; 14:357–383


Published online 27 April 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/stc.162

Prestressed active post-tensioned tendons control for


bridges under moving loads

Khaldoon A. Bani-Hani1,*,y,z and Musa R. Alawneh2,}


1
Jordan University of Science & Technology, Irbid, Jordan
2
E. Construct, FZ-LLC, Dubai, U.A.E.

SUMMARY
This paper proposes a new innovative mechanism for active bridge vibration control due to moving load
using active prestressed post-tensioned tendons. King Abdullah’s Hospital Bridge at the Jordan University
of Science and Technology is employed in this study. This bridge consists of eight spans each is 25 m
resulting in a total length of 200 m. The bridge system is treated as an elastic continuum coupled with a
single degree of freedom moving oscillator representing the moving vehicle. A linear time-varying model
used to simulate the bridge–vehicle system. The dynamic behaviour of the prestressed concrete bridge
studied for the uncontrolled case, and the effect of the prestressing force on the response of bridge–
oscillator system is illustrated. Two linear quadratic gaussian (LQG) controllers with time-varying gains
and constant gains designed to mitigate the bridge vibration due to the moving oscillator for different
velocities. The control strategy regulates the post-tensioned tendons’ force so that the system vibration
diminished. The results indicate that the proposed control mechanism is efficient for reducing the vertical
displacements of both the oscillator and the bridge by up to 83% and the vertical accelerations by up to 83
and 42%, respectively. Additionally, the robustness of the controllers are investigated and illustrated. The
results show that the proposed mechanism is robust and stable. Copyright # 2006 John Wiley & Sons, Ltd.

KEY WORDS: vibration; bridge; active control; oscillator; prestressed; moving load

1. INTRODUCTION

Recent advancements in design technology and material quality in civil engineering applications
supported the construction of lighter and more slender structures, which causes structures to be
vulnerable to dynamic loads, particularly moving loads. Large deflections and vibrations
induced by heavy and high-speed vehicles affect significantly the safety and serviceability of
bridges. Vibrations induced by moving loads may significantly increase the maximum internal
stress of bridges and deteriorate serviceability. Therefore, vibration control using special control
devices is a very efficient solution to suppress the response of bridges and, thus, to enhance the

*Correspondence to: Khaldoon A. Bani-Hani, Jordan University of Science & Technology, Irbid, Jordan.
y
E-mail: khaldoon@just.edu.jo
z
Assistant Professor.
}
Structural Engineer.

Received 27 July 2005


Revised 20 October 2005
Copyright # 2006 John Wiley & Sons, Ltd. Accepted 31 October 2005
358 K. A. BANI-HANI AND M. R. ALAWNEH

structural safety and serviceability of bridges. The dynamic response of the bridge–vehicle
system depends on the dynamic properties of the traversing vehicles and bridge, vehicle
operating speeds, surface quality of the roadway among others.
Few researchers have studied active control of bridges under moving loads. Kwon et al. [1]
analysed the behaviour of studied three span bridges under moving loads and controlled the
bridges using a tuned mass damper (TMD). The TMD is considered as a passive-type control
device. The maximum vertical displacement induced by a train is decreased by 21% and free
vibration dies out quickly. Michalopoulos et al. [2] studied passive control of bridges based on
prestressed tendon systems using two-level prestressing cable systems. Permanent (dead) loads
were relieved by the first-level prestressing cables while moving loads together with excessive
displacements were supported by second-level prestressing system. Lin and Trethewey [3]
analysed a beam subjected to moving loads using the finite element method, then controlled the
beam using a spring attached below the beam to serve as both a passive and active control
device. By controlling the extension and contraction of the spring, the active moments could be
created at the left and right control arms locations to suppress the beam vibration. Stavroulakis
et al. [4] proposed a new design concept for large bridges that is based on an integrated hanging
cable and deck–structure system with unilaterally connected blocks which were stabilized by
prestressed cables. Wang et al. [5] studied the applicability of passive TMD (PTMDs) to
suppress train-induced vibration on bridges. The results from simply supported bridges of
Taiwan High-Speed Railway (THSR) under German I.C.E., Japanese S.K.S. and French
T.G.V. trains showed that the proposed PTMD was a useful vibration control device in
reducing bridge vertical displacements, absolute acceleration, end rotations and train
accelerations during resonant speeds, as the train axle arrangement was regular. Gerco and
Santini [6] studied the effect of using rotational viscous dampers on the dynamic response of
simple beams subjected to moving loads. Dyke and Johnson [7] used an active controller to
reduce the response of bridge under moving oscillator. The active control device was placed in
parallel with a spring and dashpot elements that represent the suspension of the oscillator. The
results indicated that this control system was effective in reducing the responses of both the
oscillator and the continuum.
The problem of calculating the dynamic response of a distributed parameter system carrying
one or more travelling subsystems is crucial in several engineering applications related to
bridges, cable railways, etc. Over the years, several mathematical models have been used to
simulate the behaviour of the moving loads. In general, three types of models have been
considered in literature. The first had the vehicle modelled as a moving force producing a
dynamic response of the continuum but neglecting both the inertia of the vehicle and the
dynamics of its suspension. The second modelled the vehicle as a moving mass taking into
account the inertia of the moving subsystem and assuming infinite stiffness of the coupling
between the vehicle and the distributed system. The third model represented the vehicle as a
moving oscillator model. Pesterev and Bergman [8, 9], Pesterev et al. [10], Pesterev and Bergman
[11], and Pesterev et al. [12, 13] showed that the coupled model had definitely cope the dynamic
response of the bridge–vehicle system and significantly larger than the static response alone.
In this study, a new inventive control mechanism based on the activation of the prestressed
tendons in the prestressed reinforced concrete bridge is studied. The mechanism presents a new
technique of exploiting the post-tensioned cables to suppress the bridge vibrations due to heavy
trucks and moving loads. A side, the dynamic behaviour of the prestressed concrete bridges
under moving loads is investigated. The active tendons are regulated using a linear quadratic

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 359

Gaussian (LQG) regulator. To implement the active control method proposed in this paper,
three feedback sensors were used and located on strategic spots in the bridge deck to measure
the acceleration of the bridge Furthermore, an actuator was placed on the live end of the
prestressed tendons to generate the active prestressing force to produce the required control
force.

2. METHODOLOGY

The bridge–vehicle system under study is considered as a single beam with a rigidity equivalent
to that of the combined system of floor, stringers and main girders. Additionally, only the
fundamental modes of vibration are considered while all higher modes are neglected. Finally,
the vehicle is represented by a system with one degree of freedom only, e.g. a moving oscillator
with constant speed. The weight of the vehicle is applied at the vehicle’s mass centre rather than
at the wheels’ point of contact with the bridge floor. Material properties are assumed elastic,
isotropic and homogeneous. The deflection of the beam is considered to be a result of bending
moment effects only, the inertial properties of the beam section in rotation and beam
deformation due to shear were ignored. Such a model is referred often to as the Euler–Bernoulli
beam [14, 15].
The dynamic model of the bridge–vehicle system is developed as a vehicle traversing along the
bridge with a constant speed, V. The bridge is modelled as a simply supported Euler-Bernoulli
beam. The coupled mathematical model comprehends the dynamic behaviour of the Bernoulli-
Euler beam taking into account the passive and active prestressed cables’ force exerted on the
bridge.

2.1. Equations of motion


The prestressed concrete beam considered in this study is depicted in Figure 1, its cross-sectional
area A, flexural rigidity EI, total length L, and mass density r, with a parabolic prestressing
tendon profile, eðxÞ; given by
eðxÞ ¼ 4eo ðx2  LxÞ=L2 ð1Þ
where eo is the tendon’s maximum eccentricity at the beam mid-span. In addition, the axial
prestressing force in the tendons is equal to ðTo þ Tc Þ: The initial prestressing force, T0, designed
for static load requirements and the active prestressing force, TC ; represents the active control
force exerted on the bridge. Moreover, the total static and dynamic displacements anywhere in
the bridge at any time are denoted by yt ðx; tÞ:
In this analysis, the beam is subjected to a time-dependent distributed load, Pðx; tÞ;
perpendicular to its axis. Ignoring the inertia properties of the element in rotation and summing
the moments about point O yields

@Mt ðx; tÞ @2 yt ðx; tÞ


 þ Vt ðx; tÞ ¼ 0 ) Mt ðx; tÞ ¼ EI
@x @x2
@Vt ðx; tÞ @4 yt ðx; tÞ
) ¼ EI ð2Þ
@x @x4

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
360 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 1. The internal equilibrium of the prestressed bridge.

Next, taking force equilibrium in the y-direction yields

@Vt ðx; tÞ @yðx; tÞ @2 yt ðx; tÞ


 ðTo þ TC Þ  rA
@x @x @t2
dgðxÞ
¼ Pðx; tÞ þ ðTo þ TC Þ  rAg ð3Þ
dx

where y ¼ @yt ðx; tÞ=@x; g ¼ deðxÞ=dx and g is the gravitational acceleration. Therefore, using
results in Equations (1) and (2), Equation (3) may be written as

@4 yt ðx; tÞ @2 yt ðx; tÞ @2 yt ðx; tÞ


 EI 4
 ðTo þ TC Þ 2
 rA
@x @x @t2
de2 ðxÞ
¼ Pðx; tÞ þ ðTo þ TC Þ  rAg ð4Þ
dx2

where yt ðx; tÞ ¼ yo ðxÞ þ yðx; tÞ and represents the summation of the static and dynamic vertical
displacements. Therefore, the governing partial differential equation of motion is
 4
@ yo ðxÞ @4 yðx; tÞ
 2
@ yo ðxÞ @2 yðx; tÞ
 
EI þ þ ðTo þ TC Þ þ
@x4 @x4 @x2 @x2
@2 yðx; tÞ de2 ðxÞ
 
þ rA ¼ rAg þ Pðx; tÞ  ðTo þ TC Þ ð5Þ
@t2 dx2

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 361

Additionally, the beam equation of equilibrium for static loadings alone is given by

@4 yo ðxÞ
   2   2 
@ yo ðxÞ de ðxÞ
EI þ To ¼ rAg  To ð6Þ
@x4 @x2 dx2

and the static displacement function considering the prestressing force effect is computed from

rAgL2  8To eo 4
yo ðxÞ ¼ ðx  2Lx3 þ L3 xÞ ð7Þ
24 EI L2

Assuming a uniformly distributed mass and an unvarying stiffness along the continuum and
using the results in Equations (1)–(7) the equation of motion can be rewritten as

@4 yðx; tÞ @2 yðx; tÞ @2 yðx; tÞ


EI 4
þ ðTo þ TC Þ 2
þ rA
@x @x @t2
rAgL2  8To eo 2
 
8eo
¼ Pðx; tÞ  TC ðx  LxÞ þ 2 ð8Þ
2 EI L2 L

The response of the bridge given by Equation (8) is approximated by a linear combination and a
series expansion of the eigenfunctions as follows:
N
X
yðx; tÞ ¼ UT ðxÞqðtÞ ¼ fn ðxÞqn ðtÞ ð9Þ
n¼1

where UT ðxÞ ¼ ½f1 f2 ::: fn :::; and fn ðxÞ is the mode shape function of the nth mode of the
simply supported beam given by
sffiffiffiffiffiffiffiffiffiffi Z L
2 npx
fn ðxÞ ¼ Sin such that rA f2n ðxÞ dx ¼ 1 ð10Þ
rAL L 0

Also, qðtÞ ¼ ½q1 q2 ::: qn :::T ; and qn ðtÞ is the nth modal amplitude and is defined as
Z L
q. n ðtÞ þ o2nT qn ðtÞ ¼ fr ðxÞPðx; tÞ dx  En TC for n ¼ 1; 2; 3; . . . ; N ð11Þ
0

that is obtained by substituting Equation (9) into Equation (8), multiplying each term by fr ðxÞ;
integrating with respect to x between 0 and L, and applying the orthogonal properties of the
modes.
onT is the nth modal natural frequency of the simply supported beam considering the axial
prestressing force and given by
EI np4 ðTo þ TC Þnp2
o2nT ¼  ð12Þ
rA L rA L

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
362 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 2. The moving vehicle traversing on the prestressed bridge (bridge–vehicle model).

and En is a constant results from the integration process and given as


sffiffiffiffiffiffiffiffiffiffi
2 8eo rAgL3  8LTo eo

En ¼  ð1  cos npÞ ð13Þ
rAL npL EI n3 p3

2.2. The moving vehicle dynamics


Figure 2 shows a moving oscillator traversing along a bridge at a constant velocity replacing
the distributed load, Pðx; tÞ: The oscillator, representing the vehicle, has a total mass mv ; linear
stiffness of its suspension kv ðnatural frequency ¼ ov Þ; suspension viscous damper coefficient
cv ðdamping ratio ¼ xv Þ; vehicle velocity v, and a vertical displacement from its equilibrium
position, zðtÞ; as shown in Figure 2. Subsequently the general form of moving oscillator may be
expressed as
  
L
Pðx; tÞ ¼ HðtÞ  H t  ðmv g þ f ðtÞÞdðx  vtÞ ð14Þ
v
where Hð : Þ is the Heaviside function, dð Þ is the Dirac’s delta function and, f ðtÞ is the force
applied to the sprung mass and beam and is expressed as
f ðtÞ ¼ kv ðzðtÞ  yðx; tÞÞ þ cv ð’zðtÞ  y’ ðx; tÞÞ ð15Þ
The moving oscillator dynamics governed by equations of motion given as follows:
mv z. þ f ðtÞ ¼ 0 ) z. ¼ o2v ðyðx; tÞ  zðtÞÞ þ 2xv ov ð’yðx; tÞ  z’ ðtÞÞ ð16Þ
Substituting Equation (9) into Equation (15), the force expression is
f ðtÞ ¼ kv zðtÞ  UT ðxÞqðtÞ þ cv z’ ðtÞ  UT ðxÞ’qðtÞ
 
ð17Þ

2.3. Time-varying state-space


The derivations of the linear dynamic equation of equilibrium for the bridge–vehicle system
presented in the proceedings paragraphs are formulated in a state-space form. This is
accomplished

by first limiting the time of attention when the vehicle is traversing the bridge (i.e.
t 2 0; L=v ), and next substituting Equation (14) into Equation (11) and integrating to get
q. n ðtÞ þ o2nT qn ðtÞ ¼ fn ðvtÞðmv g þ f ðtÞÞ  En TC for n ¼ 1; 2; 3; . . . ; N ð18Þ

Substituting Equations (9) and (17) into Equation (16) and Equation (17) into Equation
(18), the coupled time-varying system is produced as below
z. ðtÞ ¼ o2v UT ðvtÞqðtÞ  o2v zðtÞ þ 2xv ov UT ðvtÞ’qðtÞ  2xv ov z’ ðtÞ ð19Þ

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 363

q. n ðtÞ ¼  o2nT qn ðtÞ þ fn ðvtÞkv zðtÞ  fn ðvtÞkv UT ðvtÞqðtÞ


þ fn ðvtÞcv z’ ðtÞ  fn ðvtÞcv UT ðvtÞ’qðtÞ þ fn ðvtÞmv g  En TC ð20Þ

Finally, defining a state vector xðtÞ ¼ ½z q z’ q’ T ; Equations (19) and (20) are written in a
time-varying state-space as
’ ¼ AðtÞxðtÞ þ BðtÞuðtÞ
xðtÞ
yðtÞ ¼ CðtÞxðtÞ þ DðtÞuðtÞþtðtÞ ð21Þ
yr ðtÞ ¼ Cr ðtÞxðtÞ þ Dr ðtÞuðtÞþtðtÞ

where uðtÞ ¼½mv g TC  is a two-dimensional vector represents the external forces, tðtÞ is the
measurement noise vector, yðtÞ ¼ ½z q z. q. T is a 2(N+1)-dimensional vector that represents
the observed response, yr ðtÞ ¼ ½ z qT is an (N+1)-dimensional vector that represents the
regulated states and AðtÞ; BðtÞ; CðtÞ; DðtÞ; Cr ðtÞ and Dr ðtÞ are time-varying state-space
coefficient matrices given as
" # " # " #
0 I 0 I 0
AðtÞ ¼ ; BðtÞ ¼ ; CðtÞ ¼ ð22Þ
A21 ðtÞ A22 ðtÞ B21 ðtÞ A21 ðtÞ A22 ðtÞ



DðtÞ ¼ ½B21 ðtÞ; Cr ðtÞ ¼ I 0 and Dr ðtÞ ¼ ½0

where
o2v o2v f1 ðvtÞ ... o2v fN ðvtÞ
2 3

 o21T þ kv ðf1 ðvtÞÞ2 kv f1 ðvtÞf2 ðvtÞ


6
7
6 kv f1 ðvtÞ kv f1 ðvtÞfN ðvtÞ 7
A21 ¼6 ð23Þ
6 7
7
6 ... ... ... ... 7
4 5
 o2NT þ kv ðfN ðvtÞÞ2


kv fN ðvtÞ kv fN ðvtÞf1 ðvtÞ ...

2xv ov 2xv ov f1 ðvtÞ ... 2xv ov fN ðvtÞ


2 3

cv ðf1 ðvtÞÞ2


6 7
6 cv f1 ðvtÞ cv f1 ðvtÞf2 ðvtÞ cv f1 ðvtÞfN ðvtÞ 7
A22 ¼6 ð24Þ
6 7
7
6 ... ... ... ... 7
4 5
cv fN ðvtÞ cv fN ðvtÞf1 ðvtÞ ... cv ðfN ðvtÞÞ2

2 3
0 0
6 7
6 f1 ðvtÞ E1 7
B21 ¼6 ð25Þ
6 7
7
6 ... ... 7
4 5
fN ðvtÞ EN

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
364 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 3. King Abdullah Hospital Bridge, Irbid, Jordan.

2.4. Verification model


To illustrate the effectiveness of the proposed control mechanism, King Abdullah Hospital
Bridge (Irbid, Jordan University of Science and Technology in Jordan) is used for verification.
The bridge under study consists of eight spans each span is 25 m long resulting in a total length
of 200 m. Each span is regarded as a simply supported beam with a cross-section as shown in
Figure 3. The bridge has four lanes, each lane is 3.7 m wide and the total width of the bridge is
19 m, supported on nine piers. The bridge is design as prestressed concrete BIV-48 box-girder
un-bonded post-tensioned parabolic tendons. Two post-tensioned tendons sets utilized: passive
tendons and active tendons. The bridge is modelled as a Bernoulli–Euler beam with equal
stiffness regardless of its complex box-type cross-section.
The bridge mass density r ¼ 2840 kg/m3, concrete Young’s modulus Ec ¼ 25 907 MPa; and
concrete compressive strength fc0 ¼ 30 MPa: Additionally, the post-tensioned steel cables’
modulus of elasticity Est ¼ 200 GPa: The total length between supports L ¼ 25 m; girder cross-
sectional area AC ¼ 1:81 m2 and moment of inertia Ic ¼ 0:2311 m4 : Three natural modes are
considered in the eigenfunction’s expansion to express the response of the bridge in terms of its
eigenmodes (Equation 9). The mass of the vehicle mv ¼ 30 tonnes and the vehicle’s spring
stiffness kv ¼ 6900 kN=m: The vehicle natural frequency, if considered as an independent
subsystem, ov ¼ 15:16 rad=s and the damping coefficient of the vehicle corresponds to
a damping ratio xv ¼ 2% in the uncoupled system.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 365

2.5. Control force design limits


The total prestressing force in the bridge serves the purpose of controlling both cracking and
deflection, and permits the member to respond essentially as if it were un-cracked. When the
member is subjected to the prestressing force ðTo þ TC Þ and the moving vehicle, the concrete
stress f1 at the top face of the member and f2 at the bottom face can be computed by
superimposing axial and bending effects:
 
ðTo þ TC Þ eo C1 Mt
f1 ¼  1 2  ð26Þ
AC r S1

 
ðTo þ TC Þ eo C2 Mt
f2 ¼  1þ 2 þ ð27Þ
AC r S2
where AC is the area of the concrete cross-section, r2 is the radius of gyration
(r2 ¼ IC =AC ¼ 0:1276 m2 ), C1 and C2 are the distance from the concrete cancroids to the top
and bottom surfaces of the member, respectively, and S1 ¼ IC =C1 ¼ 0:4622 m3 and S2 ¼
IC =C2 ¼ 0:4622 m3 are the section moduli with respect to the top and bottom surfaces of the
beam. In addition, eo is the tendon’s maximum eccentricity at beam mid-span and is equal to
0.4 m and Mt is being the total moment due to self-weight and moving load at mid-span given as
rAc gL2 mv gL
Mt ¼ þ ¼ 5778 kN m ð28Þ
8 4
The concrete stress should not exceed the tensile rupture stress or the allowable compression
stress in the concrete. Thus, to find the minimum and the maximum control force, f1 and f2 are
controlled by Equation (29) [16].
0 pffiffiffiffi
fcs ¼ 0:45fc ¼ 13:5 MPa and fts ¼ 0:5 fc0 ¼ 2:7386 MPa ð29Þ
Substituting these values in Equations (26) and (27), the maximum and minimum control forces
can be computed. Firstly, stresses at the bottom face of the beam at the mid-span should not
exceed the tensile rupture stress or the allowable compression stress given in Equation (29).
That is
 
ðTo þ TC Þ ð0:4Þð0:5Þ 5778
13:5 MPa 4  1þ þ 4 2:738 MPa ð30Þ
1:81 0:1276 0:4622
Therefore, 18334 kN 5To þ TC 5 6885 kN. Nevertheless, the top fibre stress should not exceed
the allowed compression stress for concrete or the rupture stress
 
ðTo þ TC Þ ð0:4Þð0:5Þ 5778
2:738 MPa 5  1  5  13:5 MPa ð31Þ
1:81 0:1276 0:4622
Consequently, 48 654 kN 5To þ TC 5 3184 kN (Tension). The lower value indicates a
compressive force on the tendons which is not obtainable; hence, the lower value is replaced by
zero. Therefore, the above calculations generate the lower and upper bounds of the prestressing
force: 6885 kN 4To þ TC 4 18 334 kN. Additionally, To is the required prestressing force to
balance the self-weight of the beam and calculated as follows
wselfweight L2 rAgL2
To ¼ ¼ ¼ 9848 kN ð32Þ
8e0 8e0

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
366 K. A. BANI-HANI AND M. R. ALAWNEH

Equation (32) is derived based on an equivalent upward load produced by a parabolic tendons
(To) equivalent to downward load produced by bridge’s own weight, in which all transverse
forces are cancelled [17]. For this unique balanced load condition, there are no bending
moments applied to the beam, no flexural stresses exist, but only the axial stresses produced by
the longitudinal component of the prestressing forces (axial compression force).
Subsequently, the lower and upper bounds of the control force, TC were found to be
Zero 4 TC 4 8486 kN ð33Þ
This result was considered in developing the proposed LQG controllers.

2.6. Active control force losses due to friction


The active control force in post-tensioned members is generated using unbonded post-tensioned
tendons. The tendons are anchored at one end and controlled with hydraulic actuator in the
other end. As the steel tendon slides through the duct, friction is buildup and force at the
anchored end is less than the hydraulic end. These losses are results of wobble friction, due to
unintentional misalignment, and curvature friction. When a curved tendons is used to prestress a
beam, the total tension loss in the unbonded pre-greased wire tendons due to friction (ACI Sec.
18.6.2) [18] is computed by
TC ðx; tÞ ¼ TC ðL; tÞ=eKxþmaðxÞ ð34Þ
where TC ðx; tÞ is the developed force at distance x from the active end, TC ðL; tÞ is the issued
active force at the active end, K is a wobble friction coefficient (m1), m is friction coefficient
between tendon and duct and a is the total angle change between active end and distance x from
it. Table R18.6.2 of ACI building code commentary, 318/318R-270 [18], suggests average
numerical values for K ¼ 0:0037 m1 and m ¼ 0:1: Considering these values and estimating
aðxÞ ¼ deðxÞ=dx using Equation (1), then the maximum loss at the anchored end (x ¼ L ¼ 25 m;
a ¼ 0:064 rad) is estimated to be 9% of the active force. Therefore, the control force issued by
the actuator has to be increased by 9% to substitute for the losses induced by friction.
Immediate losses due to anchorage slip and concrete elastic shortening, and time-dependent
losses due creep, shrinkage of concrete and steel relaxation can be developed, however, these
losses are either initially or with time lost while the major losses induced when the active tendons
are in operation are generated by the friction [17].

2.7. Response of the uncontrolled bridge


The resulting time-varying linear state-space model given in Equations (21) and (22) was
encoded in a SIMULINK [19] model using MATLAB [20] software. The mathematical model
served to simulate the response of the prestressed bridge under the moving vehicle (oscillator).
The uncontrolled response is referred to when TC is set to zero. This response is compared to the
controlled response in order to examine the feasibility and effectiveness of the proposed control
mechanism (Figure 4).
The axial compression force exerted by the prestressing tendons reduces the natural frequency
of the prestressed beam as may be seen in Equation (12). However, the cost of natural frequency
reduction in dynamic vibration with prestressed tendons is much less than the cost of static
deflection without the prestressing tendons. Figure 5 shows the variation of the first
fundamental natural frequency of the prestressed beam as the control force changes. It was

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 367

Figure 4. Controller configuration for bridge–vehicle model.

Figure 5. Effect of axial compressive force on the bridge dynamic response.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
368 K. A. BANI-HANI AND M. R. ALAWNEH

found that the natural frequency was reduced by approximately 2%. This result is reflected on
the response of the active prestressed bridge in the dynamic analysis as shown in Figure 5. In
short, the effect of axial compression force was that of ‘stiffness degradation’, which decreased
the natural frequencies of the beam. Consequently, the beam became more flexible, and the
deflection of the bridge increased. Additionally, it was noted that the vehicle speed affected
the response of bridge–vehicle response significantly. For example, the vertical displacement of
the vehicle and the bridge were increased by 40 and 17%, respectively, as the vehicle’s speed
varied from 10 to 35 m/s. Also, the vertical acceleration of the vehicle and the bridge were
increased by 83 and 73.5%, respectively.

3. CONTROL DESIGN

As stated earlier, the prestressed tendons exert variable forces as the vehicle transverses the
bridge. To make optimal use of the tendons, the output force has to follow optimal path to
minimize both vehicle vertical motion as well as the bridge vertical displacement. Two optimal
linear quadratic regulators (LQR) were designed and constructed as a prelude to control the
bridge–vehicle system. The first controller calculates the optimal control force TC at any time
using time-varying gains based on a full state feedbacks that is, xðtÞ ¼ ½z q1 q2 q3 z’ q’ 1 q’ 2 q’ 3 T :
The controller path was determined using a performance index that weighted the regulated
output vector, yr(t), and given as:
Z t 
# ¼ lim 1E ðCr ðtÞxr þ Dr ðtÞTC ÞT Q ðCr ðtÞxr þ Dr ðtÞTC Þ þ R½TC 2 dt

JðtÞ ð35Þ
t!1 t 0

Next, a time-varying control gain KTV ðtÞ is computed (for t ¼ 0; :::; L=V), such that JðtÞ # is
minimum and E [.] is the steady-state covariance. The solution of this optimization problem
results in the optimal time-varying controller gain KTV ðtÞ: Thus, in this case, it is assumed that
the controller knows the location of the oscillator at any time. Basically, KTV ðtÞ is the full state
feedback gain matrix for the stochastic optimal controller for the time-varying state-space. The
calculation of the gain matrix KTV ðtÞ was accomplished using MATLAB [20] code that utilized
the lqry.m routine within the control toolbox. The controller was designed by weighing the
vertical displacement of the vehicle, z, and three generalized coordinates associated with the
first, second and third vibration modes of the bridge, namely, q1 ; q2 and q3 : The weighing
matrices Q ¼ diag([0.1 5 5 5])*1  107 and R ¼ 1  106 were used in Equation (35). The second
LQR controller was developed using a constant gain matrix KCG such that a time independent
performance index Jðt# ¼ L=2VÞ is minimum, which is the case when the vehicle is at bridge mid-
span. Yet again, the same weighing matrices were considered in solving the optimization
problem in Equation (35). Figure 6 shows the time-varying gains variation as function of the
time for a vehicle travelling at speed V ¼ 25 m=s: The constant gain for the LQR controller was
estimated as KCG=[66408062, 6046952, 0.044015, 20464, 2520673, 193034, 0.002534,
886.23] that are the values of the time-varying gain at t ¼ 0:5 s: The active prestressed tendons
were controlled by the two optimal controllers given as follows:

TC ðtÞ ¼ KTV ðtÞ xðtÞ for time-varying gain matrix


ð36Þ
TC ðtÞ ¼ KCG xðtÞ for constant gain matrix

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 369

Figure 6. Time-varying gain values variation as function of vehicle travelling time.

Next, an H2/LQG control algorithm based on acceleration feedback was used to develop two
LQG controllers, specifically, the time-varying gain linear quadratic Gaussian controller and the
constant gain linear quadratic Gaussian controller. These accelerations are measured with
sensors located at certain strategic points on the bridge. The prescribed controllers’ paths were
designed using the LQG control method with a Kalman estimator.
Z t 
J# K ¼ lim E ðxðtÞ  xðtÞÞðxðtÞ
# # T
 xðtÞÞ ð37Þ
t!1 0

where xðtÞ
# is the Kalman filter estimate of the state vector xðtÞ: An optimal solution for
Equation (37) is the Kalman filter optimal estimator, x;
# and is given by

’# ¼ AðtÞ xðtÞ
xðtÞ # þ BðtÞ TC ðtÞ þ Lð.qðtÞ  CðtÞ xðtÞ
#  DðtÞ TC ðtÞÞ ð38Þ

where L is Kalman gain filter and was found using MATLAB routine lqew.m within the Control
Toolbox. Further, the measurement noises were assumed to be uniformly distributed, statically
independent Gaussian white noise processes. Then, the control force is expressed as

TC ðtÞ ¼ KTV ðtÞ xðtÞ


# for time-varying gain matrix
ð39Þ
TC ðtÞ ¼ KCG xðtÞ
# for constant gain matrix

Once the gain matrices are computed, the controllers described above were formulated in a
continuous state-space form as follows:

’ ¼ AC ðtÞhðtÞ þ BC ðtÞyC ðtÞ


SðtÞ
ð40Þ
TC ðtÞ ¼ CC ðtÞSðtÞ þ DC ðtÞyC ðtÞ

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
370 K. A. BANI-HANI AND M. R. ALAWNEH

Here, AC ðtÞ; BC ðtÞ; CC ðtÞ and DC ðtÞ are the time-varying state-space control coefficients and given
as follows:

AC ðtÞ¼ AðtÞBðtÞKTV ðtÞ  L½CðtÞ  DðtÞKTV ðtÞ


ð41Þ
BC ðtÞ ¼ L; CC ðtÞ ¼ KTV ðtÞ; DC ðtÞ ¼ 0

Additionally, the constant gain LQG was developed in similar way such that Equation (40) and
(41) were developed using constant state-space values as well as constant performance index in
addition to constant gain matrix.
The optimal LQG controller uses three feedback sensors representing the first three
generalized coordinate accelerations q. 1 ; q. 2 ; and q. 3 : These feedback signals are not directly
available for measurement. Therefore, to find out the control action, the algorithm requires an
estimate of the states at any time. To obtain the three feedback values mentioned above, three
accelerometers (sensors) located at strategic points on the bridge deck were utilized. Figure 4,
shows the likely implementation of bridge control. The sensors located along the bridge deck
measured the vertical acceleration at certain points. These accelerations were related to the
generalized coordinates q. 1 ; q. 2 and q. 3 in the following fashion:
2 3 2 32 3
y. 1 ðL=4; tÞ f1 ðL=4Þ f2 ðL=4Þ f3 ðL=4Þ q. 1 ðtÞ
6 7 6 76 7
6 y. 2 ðL=2; tÞ 7 ¼ 6 f1 ðL=2Þ f2 ðL=2Þ f3 ðL=2Þ 7
54 q. 2 ðtÞ 5 ð42Þ
6 7
4 5 4
y. 3 ð3L=4; tÞ f1 ð3L=4Þ f2 ð3L=4Þ f3 ð3L=4Þ q. 3 ðtÞ
As a result, the following transformation matrix was obtained to determine the generalized
coordinates, q. 1 ; q. 2 ; q. 3 from acceleration feedback at the three selected points (Figure 4):
2 3 2 32 3
q. 1 90:9 128:2 90:9 y. 1 ðL=4; tÞ
6 7 6 76 7
6 q. 2 7 ¼ 6 128:2 128:2 7
0 54 y. 2 ðL=2; tÞ 5 ð43Þ
6 7
4 5 4
q. 3 90:9 128:2 90:0 y. 3 ð3L=4; tÞ
The control displacement of the active tendons as well as the actuator stroke displacement can
be estimated based on the developed strains in the active tendons ignoring the shrinkage and
creep strains. In general, the axial strain developed in the active tendons’ centre line considering
the tendons’ friction loss is given as
eðxÞ2
 
TC ðx; tÞ Mv ðx; tÞeðxÞ
est ðx; tÞ ¼ 1 þ 2 þ ð44Þ
EC AC r E C IC
where Mv ðx;tÞ ¼mv gdðx  vtÞ ðLx  x2 Þ=2 is the moment function
RL produced by the moving
vehicle. Hence, the tendon displacement is defined as xT ðtÞ¼ 0 est ðx; tÞ dx;

therefore
the total
extension in the active tendon from the active end to the dead end for t 2 0; L=v ; is integrated
and given as
xT ðtÞ ¼ ð8:4978TC ðtÞ  6292v4 t4 þ 31460v3 t3  393245v2 t2 Þ  1010 ð45Þ
Consequently, the required energy at any time to move the anchorage is estimated as follows:
Z tZ L
EnðtÞ ¼ TC ðx;tÞ est ðx;tÞ dx dt ð46Þ
0 0

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 371

Figure 7. SIMULINK model for both time-varying LQG controller and constant gain LQG controller.

Finally, Figure 7 illustrates a SIMULINK model for both the time-varying and constant gains
LQG controllers.

4. ANALYTICAL RESULTS

4.1. Performance assessment for the active prestressed tendons


The controllers designed in the previous section were intended to alleviate the vibration response
of both the vehicle and the bridge under study. To achieve this, new mechanism utilizing active
prestressed tendons was employed to mitigate both the bridge vibration response as well as the
vehicle response. The effectiveness of the proposed mechanism was investigated using the linear
quadratic regulators LQR; time-varying gain and constant gain controllers, which were
developed as a preface for two other acceleration feedback linear quadratic Gaussian LQG
controllers. Results of LQR and LQG controllers, when the vehicle is traversing at a constant
speed of 25 m/s, are presented in part as shown in Figures 8 and 9. Figure 8 presents a
comparison of the controlled and uncontrolled response of the bridge. It is apparent from
Figure 8 that the vertical deflection of the bridge was reduced by about 89 and 88% for the time-
varying LQR and the constant gain LQR, respectively. Similarly, the response was reduced by
about 87 and 85% for the time-varying LQG and the constant gain LQG, respectively. It is clear
that the performance differences in the LQR and LQG controllers were very small.
Consequently, the LQG controllers are of the main focus in this study due to their practicality
and implementation simplicity. The results of the response of the controlled vehicle are
presented in Figure 9, which shows that the vertical response of the vehicle is compared to the

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
372 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 8. Comparisons of the vibration response of the uncontrolled and controlled bridge for a moving
vehicle with speed of 25 m/s, using time-varying and constant gains LQR and LQG controllers.

uncontrolled and the controlled case for both the LQR and the LQG. Moreover, Figure 9
reveals that the maximum control forces exerted by the active tendons, TCmax ; was 3686 kN, and
compares results for the different controllers mentioned above. The difference between the
control forces generated by LQR or LQG is small and in the range of 4%. It was also noted that
the required control force is varying smoothly with almost constant acceleration. Additionally,
the stroke displacements and the applied force of the active actuator are shown in Figure 9. Peak
values of the stroke are about 4.41 mm for constant gain LQG and about 4.55 mm for time-
varying LQG and maximum exerted energy is about 7.2 kN m.
The effectiveness of the active prestressed tendons is illustrated numerically using time-
varying and constant gain LQG controllers. The results are presented in Tables I and II. Table I

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 373

Figure 9. Comparisons of the control forces and the energy and actuator stroke history for the
different controllers, in addition the vertical vibration response of the uncontrolled and controlled
moving vehicle with speed of 25 m/s is compared.

compares the peak values of the vertical displacements and vertical accelerations in the
uncontrolled and the controlled bridge–vehicle system and outline the corresponding peak
values of the active tendon forces and the actuator stroke displacement for the moving vehicle at
various speeds ranging from 10 to 35 m/s. It is clear from Table I that the constant gain LQG
controller was effective in peak response reduction. For example, the vertical displacements and
accelerations of the vehicle decreased by on average value of 83.5% and, on average, the bridge
vertical displacement decreased by 83.5% and the vertical acceleration decreased by 42% for the
selected speed range. Similarly, Table II presents the same comparative results for the controlled
and uncontrolled peak values of the vertical response of the vehicle–bridge when traveling at
various speeds of 10–35 m/s for the time-varying LQG controller. Results demonstrate that the

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
Copyright # 2006 John Wiley & Sons, Ltd.

374
Table I. Comparisons of the controlled and uncontrolled peak response for the bridge–vehicle system using the constant gain controller LQG.
Bridge vertical Bridge vertical Vehicle vertical Vehicle vertical
displacement acceleration displacement acceleration
Vehicle (mm) (m/s2) (mm) (m/s2) Stroke
speed Control displs.
(m/s) Uncont. Controlled Uncont. Controlled Uncont. Controlled Uncont. Controlled force (kN) (mm)
10 18.5 4.0 (78%) 0.6166 0.4111 (33%) 19.5 3.9 (80%) 0.4510 0.0415 (91%) 3253 4.30
15 19.9 3.5 (82%) 0.8699 0.5162 (41%) 19.5 3.5 (82%) 0.6868 0.1143 (83%) 3355 4.39
20 19.3 3.3 (83%) 1.2350 0.6855 (45%) 19.3 3.9 (80%) 1.1330 0.1946 (83%) 3320 4.36

K. A. BANI-HANI AND M. R. ALAWNEH


25 21.1 3.1 (85%) 1.4121 0.8524 (40%) 24.8 4.0 (84%) 1.6100 0.3344 (79%) 3395 4.41
30 22.0 3.0 (86%) 1.7801 1.0164 (43%) 29.4 3.8 (87%) 2.1260 0.4521 (79%) 3503 4.51
35 22.2 3.0 (87%) 2.3273 1.1758 (50%) 33.1 4.0 (88%) 2.6971 0.3792 (86%) 3577 4.57
Avg. 83.5 42 83.5 83.5
reduction (%)

Table II. Comparisons of the controlled and uncontrolled peak response for the bridge–vehicle system using the time
varying gain controller LQG.
Struct. Control Health Monit. 2007; 14:357–383

Bridge vertical Bridge vertical Vehicle vertical Vehicle vertical


displacement acceleration displacement acceleration
Vehicle (mm) (m/s2) (mm) (m/s2) Stroke
speed Control displs.
(m/s) Uncont. Controlled Uncont. Controlled Uncont. Controlled Uncont. Controlled force (kN) (mm)
10 18.5 3.9 (79%) 0.6166 0.3777 (39%) 19.5 4.0 (80%) 0.4510 0.0706 (84%) 3290 5.00
15 19.9 3.2 (84%) 0.8699 0.5315 (39%) 19.5 3.6 (77%) 0.6868 0.1510 (78%) 3484 4.49
20 19.3 3.0 (85%) 1.2350 0.6832 (45%) 19.3 3.9 (80%) 1.1330 0.2553 (77%) 3532 4.53
25 21.1 2.8 (87%) 1.4121 0.8490 (40%) 24.8 3.5 (86%) 1.6100 0.2515 (84%) 3548 4.55
30 22.0 2.8 (88%) 1.7801 1.0112 (43%) 29.4 3.1 (89%) 2.1260 0.2895 (86%) 3629 4.62
35 22.2 2.6 (88%) 2.3273 1.1678 (50%) 33.1 3.2 (90%) 2.6971 0.2843 (89%) 3820 4.76
Avg. 84.5 42.7 83.7 83
reduction (%)
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 375

Figure 10. Comparisons of the vertical acceleration of the uncontrolled and controlled bridge for a moving
vehicle with speed of 25 m/s, using time-varying and constant gains LQG controllers.

time-varying LQG controller reduced the vehicle vertical displacement by 83.7% and the
vertical acceleration by 83% on the average for the selected speeds. Moreover, comparison
between the bridge peak values is illustrated for the time-varying LQG. In this case, the peak
values were reduced by 84.5 and 42.7%, respectively. The maximum control force as well as the
actuator stroke peaks for the LQG controllers are shown in Tables I and II and the maximum
force is within the limits discussed in Section 4.
In general, the results show that the constant gain LQG and the time-varying LQG
controllers were successful in bridge–vehicle vibration mitigation as well. Again, it is clear that
the constant gain controller performed as well as the time-varying controller with insignificant
differences. Based on this observation it was decided to adopt the constant gain LQG as a
general controller throughout this study.
Additional demonstration is presented in Figure 10 that displays the vertical acceleration of
the uncontrolled and controlled bridge for a moving vehicle with a constant speed of 25 m/s.
Figure 10 shows that the constant gain LQG controller reduced the maximum vertical
acceleration to 0.8524 m/s2 at distance of 5.25 m from the beam support when the vehicle is at
2.25 m from the support, compared to the maximum uncontrolled response of 1.4121 m/s2 at
distance equal to 11.5 m and traveling time equal to 0.27 s. On the other hand, the time-varying
LQG controller decreased the peak vertical acceleration to 0.849 m/s2 at a distance of 5.25 m
when the vehicle is 2.25 m away from the starting support.
Additional results are depicted in Figures 11–13 where the controlled responses of the moving
vehicle and the bridge are outlined for two vehicle speeds, 15 and 35 m/s as the constant gain
LQG controller operates on the active prestressed tendons. Figure 11 shows that the peak values
of the vehicle vertical displacement decreased by 82% for V ¼ 15 m=s and by 88% for
V ¼ 35 m=s: Similarly, the peak values of the vehicle vertical acceleration were reduced by 83
and 86% for V ¼ 15 and 35, respectively. Figure 12 shows that the bridge vertical response for
V ¼ 15 m=s was reduced, by 82.4 and 40.7%, for the displacement and acceleration,
respectively. Similar comparison is depicted in Figure 13 for V ¼ 35 m=s revealing peaks
reductions by 86.5 and 49.5%. It is noteworthy that the constant gain LQG was developed for

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
376 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 11. Comparisons of the vertical response of the uncontrolled and controlled moving vehicle with
speeds of 15 and 35 m/s, using constant gains LQG controllers computed for V ¼ 25 m=s.

the case of vehicle speed V ¼ 25 m=s: Numerical comparison is also presented in Tables I and II
as mentioned earlier.

4.2. Robustness of the active prestressed tendons


In order to explore the robustness of the active prestressed mechanism, a number of analyses
were performed and discussed. In the first place, the bridge–vehicle model uncertainty was
considered for two conditions: the flexural rigidity, EI, and the mass density, r, of the bridge by
20%: Those situations are obtained by multiplying the original values of the EI and r by 1.2
or 0.8, and denoting the results by ‘D(EI, r) =20 %’. Additionally, uncertainty in the
controller state-space coefficients (AC, BC, CC and DC) was considered by adjusting their values
by 15 % and referring to the results by ‘(AC, BC, CC, DC)=15 %’. Likewise, effects of time-
delay as well as high-level measurement noises were examined. The time-delay considered was
0.01 s and the measurement noises were modelled as Gaussian rectangular pulse processes. The
pulse width was 0.01 s and a two-sided spectral density of 0.1, 0.25 and 0.25 m2Hz/s3 for the
generalized coordinates q. 1 ; q. 2 and q. 3 to give noise level up to 10; 15; 15 m=s2 ;
respectively. The controllers were experimented when some feedback sensors failed to send the
correct signal and transmitted measurement noise instead. Finally, the stability of the active
tendons for friction loss over/under-estimation is evaluated.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 377

Figure 12. Comparisons of the vertical response of the uncontrolled and controlled bridge for a moving
vehicle with speeds of 15 m/s, using constant gains LQG controllers computed for V ¼ 25 m=s.

The uncertainty in flexural rigidity and the mass density for the two conditions ‘D(EI, r) =
20 %’, were obtained and response analyses were carried out using the time-varying
SIMULINK model. The peak values response quantities for the above-mentioned robustness
analyses were calculated for the constant gain LQG controller with three velocity cases represent
the expected mean velocity, V ¼ 25 m=s; lower (trucks) velocity, 15 m/s, and upper (fast car)
velocity, 35 m/s, and presented in Table III. Results in Table III reports the vertical responses of
the vehicle-bridge system, zmax ; ymax ðx; tÞ; y. max ðx; tÞ and the control forces for the robustness
analysis cases discussed above. As seen, the peak response quantities are susceptible to the
flexural rigidity and mass density uncertainties whereas peak response values were slightly
affected by controller uncertainty. Yet, in all cases, the controller was robust within the scope of
its design. Equally, the constant gain LQG controller is sensitive to the case of high sensors
noise combined to a time-delay. In comparison to the response of the controlled system, the
peak values of the bridge–vehicle system for the D(AC, BC, CC, DC)=15 % cases, were slightly
affected. Moreover, the peak values of the vehicle vertical displacement, zmax ; bridge vertical
displacement, ymax ðx; tÞ; and vertical acceleration, y. max ðx; tÞ; for the D(EI, r) =20 %; increased,

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
378 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 13. Comparisons of the vertical response of the uncontrolled and controlled bridge for a moving
vehicle with speeds of 35 m/s, using constant gains LQG controllers computed for V ¼ 25 m=s:

respectively, by 17, 14 and 27%, and for the D(EI, r) =þ20 %; decreased by 11.4%, 11.4%
and 18.1% for V ¼ 15 m=s: Similar percentage can be computed using results in Table III for
V ¼ 25 and 35 m/s. Furthermore, the sensor noise combined to the time-delay has affected the
peak values by 3, 15 and +16%, for V ¼ 25 m=s; respectively, again results outline other
values for V ¼ 15 and 35 m/s. In general, it is advised not to overestimate the EI and r of the
bridge when designing the controller. On the other hand, the peak values of the vehicle–bridge
system: z; ymax ðx; tÞ and y. max ðx; tÞ; were affected slightly to severely as some sensors have failed
to operate as required. Results in Table III show that the peak response values are insensitive to
the failure of sensors 2 and 3. However, failure of sensor 1 had a catastrophic effect on the
controller performance. As shown in Table III, the controller was paralysed and did not
response as demanded but it issued about 0.1% of the active force power. These results
established the significance of the mid-span sensor that requires extra emphasis on its robustness
and design. Therefore, it is recommended to install additional sensor in the mid-span in case of
first one failure, to ensure control system effectiveness. Additionally, Table III discloses the
bridge–vehicle response when the control force was amplified and reduced by 50% for friction

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
Copyright # 2006 John Wiley & Sons, Ltd.

TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS


Table III. Performance of the constant gain LQG controllers with severe changes in controller’s coefficients or vehicle–bridge modelling or high
level of sensor noises and time delay, and sensors’ failure (robustness analysis) for V ¼ 15; 25 and 35 m/s.
Vehicle–bridge Vehicle–bridge Vehicle–bridge
peak values peak values peak values
V ¼ 15 m=s V ¼ 25 m=s V ¼ 35 m=s
Analysis case Zmax ymax y. max Tcmax zmax ymax y. max Tcmax zmax ymax y. max Tcmax
(robustness analysis) (mm) (mm) (m/s2) (kN) (mm) (mm) (m/s2) (kN) (mm) (mm) (m/s2) (kN)
Uncontrolled 19.5 19.9 0.8699 NA 24.8 21.1 1.4121 NA 33.1 22.2 2.3273 NA
Controlled (constant LQG) 3.5 3.5 0.5162 3355 4.0 3.1 0.8524 3395 4.0 3.0 1.1758 3577
Sensor #1 failed 19.4 19.9 0.7684 4.14 24.8 21.0 1.2524 4.42 33.1 22.2 1.9489 5.34
Sensor #2 and #3 failed 3.2 2.9 0.5144 3429 3.9 3.0 0.8506 3442 3.8 2.8 1.1734 3611
D(AC, BC, CC, DC) = +15% 3.0 2.8 0.5222 3523 3.5 4.1 0.8250 3184 3.2 2.4 1.1883 3700
D (AC, BC, CC, DC) = 15% 4.7 4.6 0.5355 4190 4.7 2.5 0.8618 3531 5.3 4.0 1.1386 3378
Struct. Control Health Monit. 2007; 14:357–383

Sensors’ noises and time delay 3.4 3.2 0.6013 3390 3.9 3.0 1.0075 3417 4.0 2.8 1.4018 3641
DEI = +20%, Dr = +20% 3.1 3.1 0.4227 3338 3.4 2.8 0.7007 3391 3.5 2.6 0.9678 3562
DEI = 20%, Dr = 20% 4.1 4.0 0.6569 3357 4.8 3.6 1.0807 3382 4.6 3.4 1.4886 3579
Control force = 150%  TC 2.9 2.7 0.5191 3535 3.5 2.5 0.8565 3536 3.2 2.3 1.1806 3705
Control force = 50%  TC 5.7 5.5 0.5805 2907 5.3 4.9 0.8018 3034 6.3 4.8 1.1084 3244

379
380 K. A. BANI-HANI AND M. R. ALAWNEH

Figure 14. Comparison between the bridge responses when the constant gain LQG controller has been
modified by 15% of its computed values and when some sensors failed to report the correct feedbacks
for vehicle traversing at 25 m/s (robustness analysis).

loss. Results proved that the control force was stable and effective with friction loss
miscalculation or variation for any reason.
For more demonstrations of the system robustness, Figure 14 reveals the effects of 15 %
uncertainty on controller state-space coefficients (D(AC, BC, CC, DC)=15 %) for V ¼ 25 m/s.
It may be readily seen that the performance of the LQG degraded for the bridge mid-span
vertical displacement but faintly affected the corresponding accelerations. Furthermore,
Figure 14 presents selected comparisons of the responses for the bridge when the mid-span
sensor failed and when sensors 2 and 3 failed. It is clear that the controller performance
deteriorated when the mid-span sensor stopped working but unaffected when sensors 2 and 3
failed. However, the controller performance maintained its stability, and the worst that could
happen was that the controller would stop working rather than excessively destructs the bridge.
Finally, Figure 15 layouts a comparison between the LQG controller with uncertainty in EI and
r by D(EI, r) ¼ 20% for the mean velocity, V ¼ 25 m=s: Figure 15 shows how the response of
the vehicle–bridge system was affected while the vehicle was in motion.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 381

Figure 15. Comparison between the vehicle–bridge responses when the constant gain LQG controller is
controlling a modified flexural rigidity as well as mass density by 20% of its computed values for
vehicle velocity equal to 25 m/s (robustness analysis).

The controller stability was ensured by controlling the closed-loop eigen values of the time-
varying controller state-space given in Equations (40) and (41). This was achieved by ensuring
that the real parts of the eigenvalues of the time-varying matrix Ac(t) is always less than zero
[21].

5. CONCLUSIONS

Actively controlled post-tensioned un-bonded prestressed tendons mechanism was proposed to


mitigate the dynamic response of a prestressed concrete bridge under a moving oscillator

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
382 K. A. BANI-HANI AND M. R. ALAWNEH

(vehicle). Results have indeed shown that the axial compression force decreased the natural
frequency of the system while the dynamic response slightly increased for both the bridge and
vehicle, however, the pluses of prestressed tendons in static response reduction exceed the
minuses of the slight increase in dynamic response. The performance of the controllers were
considered and outlined. The results indicated that this control system is equally effective for
reducing the vibrations of both the vehicle and the bridge. Additionally, results revealed that the
time-varying gain LQG performed slightly better than the constant gain LQG. However, both
controllers were quite successful in reducing the bridge–vehicle vibration response, thus, the
constant gain controller was deemed more practical and simpler. High-level measurements
noises, computation time-delay, model uncertainties, controller uncertainties and feedback
sensors’ failure, are all severe conditions considered in the robustness analysis for the LQG
controller. Furthermore, the study indicated that the LQG controller is robust within the
bounds of its scope but severely deteriorated for the mid-span sensor failure demonstrating the
importance of the mid-span sensor. Whatever the case may be, the controller force was observed
not to violate the allowable stresses for the concrete bridge which is reflected on the control
force limits.
Moreover, the results demonstrated the reliability and stability of the tendons in the sense of
the optimal control design as well as the significant attractive feature of using acceleration
feedbacks, which constitutes a practical, reliable and efficient control methodology. Finally, the
results of this study illustrate the potential strength of the use of active prestressed tendons for
vibration mitigation of prestressed reinforced concrete bridges under moving load.

REFERENCES
1. Kwon H-C, Kim M-C, Lee I-W. Vibration control of bridges under moving loads. Computers and Structures 1998;
66(4):473–480.
2. Michalopoulos A, Stavroulakis GE, Zacharenakis EC, Panagiotopoulos PD. A prestressed tendon based passive
control system for bridges. Computer and Structures 1997; 63(6):1165–1175.
3. Lin Y-H, Trethewey MW. Active vibration suppression of beam structures subjected to moving loads: a feasibility
study using finite elements. Journal of Sound and Vibration 1993; 166(7):383–395.
4. Stavroulakis GE, Michalopoulos A, Panagiotopoulos PD, Zacharenakis EC. A multiblock unilateral concept for
passive Control of Prestressed Bridges. Structural Multidisciplinary Optimization 2000; 19(3):225–236.
5. Wang JF, Lin CC, Chen BL. Vibration suppression for high-speed railway bridges using tuned mass dampers.
International Journal of Solid Structures 2003; 40(10):465–491.
6. Greco A, Santini A. Dynamic response of a flexural non-classically damped continuous beam under moving
loadings. Computers and structures 2002; 80(3):1945–1953.
7. Dyke SJ, Johnson SM. Active control of a moving oscillator on an elastic continuum. Proceedings of the Structural
Safety and Reliability Conference, Newport Beach, CA, U.S.A., 17–22 June 2001.
8. Pesterev AV, Bergman LA. Response of elastic continuum carrying moving linear oscillator. Journal of Engineering
Mechanics (ASCE) 1997; 123(8):878–884.
9. Pesterev AV, Bergman LA. Vibration of elastic continuum carrying acceleration oscillator. Journal of Engineering
Mechanics (ASCE) 1997; 123(8):886–889.
10. Pesterev AV, Yang B, Bergman LA, Tan CA. Response and stress calculations of an elastic continuum carrying
multiple moving oscillators. Proceedings of the International Conference on Advances in Structural Dynamics, Hong
Kong, December 2000, 13–15.
11. Pesterev AV, Bergman LA. An improved series expansion of the solution to the moving oscillator problem. Journal
of Vibration and Acoustics (ASME) 2000; 122(1):54–61.
12. Pesterev AV, Yang B, Bergman LA, Tan CA. Response of elastic continuum carrying multiple moving oscillators.
Journal of Engineering Mechanics (ASCE) 2001; 127(3):260–265.
13. Pesterev AV, Tan CA, Bergman LA. A new method for calculating bending moment and shear force in moving load
problems. Journal of Applied Mechanics 2001; 68(3):252–259.
14. Clough RW, Penzien J. Dynamics of Structures. McGraw Hill: New York, 1993.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383
TENDONS CONTROL FOR BRIDGES UNDER MOVING LOADS 383

15. Chopra AK. Dynamics of Structures: Theory and Applications to Earthquake Engineering. Prentice-Hall: New Jersey,
2002.
16. AASHTO LRFD. Bridge Design Specifications: Si Units. American Association of State Highway, 2003.
17. Nilson AH. Design of Prestressed Concrete. Wiley: New York, 1987.
18. ACI Committee. ACI 318-270, Building Code Requirements for Reinforced Concrete. American Concrete Institute,
2002.
19. The Math works Inc. SIMULINK, Natick, MA, 2001.
20. The Math works Inc. MATLAB, Natick, MA, 2001.
21. Vegate JV de. Feedback Control Systems (3rd edn). Prentice-Hall: Englewood Cliffs, NJ, 1994.

Copyright # 2006 John Wiley & Sons, Ltd. Struct. Control Health Monit. 2007; 14:357–383

You might also like