Вы находитесь на странице: 1из 99

FIBER-OPTIC

SENSORS FOR
INFRASTRUCTURE
HEALTH
MONITORING,
VOLUME II
FIBER-OPTIC
SENSORS FOR
INFRASTRUCTURE
HEALTH
MONITORING,
VOLUME II
Methodology and Case Studies

ZHISHEN WU
JIAN ZHANG
MOHAMMAD NOORI
Fiber-Optic Sensors for Infrastructure Health Monitoring, Volume II:
Methodology and Case Studies

Copyright © Momentum Press®, LLC, 2019.

All rights reserved. No part of this publication may be reproduced, stored


in a retrieval system, or transmitted in any form or by any means—­
electronic, mechanical, photocopy, recording, or any other—except for
brief quotations, not to exceed 250 words, without the prior permission
of the publisher.

First published in 2019 by


Momentum Press®, LLC
222 East 46th Street, New York, NY 10017
www.momentumpress.net

ISBN-13: 978-1-94561-222-0 (print)


ISBN-13: 978-1-94561-223-7 (e-book)

Momentum Press Sustainable Structural Systems Collection

Collection ISSN: 2376-5119


Collection eISSN: 2376-5127

DOI: 10.5643/9781945612237

Cover and interior design by S4Carlisle Publishing Services Private Ltd.,


Chennai, India

First edition: 2019

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


Abstract

Over the past two decades, extensive research has been conducted on the ap-
plication of fiber-optic sensors (FOSs) in structural health monitoring (SHM).
In Volume 1 of this book a long-gauge sensing technique for incorporating a
proposed areawise sensing, developed by the authors, was introduced. High
precision and good durability of the long-gauge sensors were also demon-
strated via technical improvements that further enable the applications of opti-
cal fiber sensors and carbon fiber sensors. In Volume 2, based on the merits of
the long-gauge sensors, the methods that have been developed for processing
areawise distributed monitoring data for structural identification are intro-
duced. A discussion follows on how those methods are capable of performing
a rich recognition of local and global structural parameters including structural
deflections, dynamic characteristics, damages, and loads. Also presented is a
three-level method of structural performance evaluation that utilizes monitor-
ing data and identified results.

KEYWORDS

area-wise distributed sensing/monitoring; damage detection; fiber-optic


sensors; infrastructure safety; long-gauge sensors; strain mode theory;
structural health monitoring; structural performance and life prediction
CONTENTS

Acknolwedgments ix
Chapter 1 Structural Identification and Damage Detection
Based on Macro-Strain Measurements 1
Chapter 2 Area Distributed Monitoring-Based Structural
Performance Evaluation 83
Chapter 3 Concluding Remarks and Future Work 123
About the Authors 127
Index 129
Acknowledgments

The authors would like to express their sincere appreciation to the fol-
lowing colleagues for their contributions: Dr. Huang Huang, Dr. Caiqian
Yang, Dr. Suzhen Li, Dr. Wan Hong, Dr. Yongsheng Tang, Dr. Sheng Shen,
and Dr. Zhang Hao. They also express their special thanks to their families
for their support. Dr. Mohammad Noori records his sincere gratitude to his
wife, Nahid, without whose unconditional support, sacrifice, and encour-
agement, throughout their many years of companionship, endeavors such
as contributions to writing this book would not have been possible.
CHAPTER 1

Structural Identification
and Damage Detection
Based on Macro-Strain
Measurements

During the process of structural health monitoring, the sensors installed


in the structure send the data collected by the data collecting (acquisition)
system to the data center. Therefore, for the performance assessment and
prediction of the structure’s behavior, it is essential to identify all state pa-
rameters, damage conditions, traffic loads, etc., from the monitoring data.
As discussed earlier, local sensing techniques and devices such as strain
meters, crack meter, corrosion apparatus, etc., are too local, and therefore,
and it is very difficult to identify the damages using these types of sen-
sors and the narrowly focused data. The global sensing techniques, such
as accelerometer, Global Positioning System (GPS), displacement meter,
inclinometer, etc., are too macroscopic, which makes it hard to identify
damages via the global information they provide. Based on what we dis-
cussed and the concepts introduced in Chapter 3 of Volume I, however,
structural damages can best be identified through the distributed response
of all critical or key areas. Following the notion of distributed sensing
and critical areas, we introduce the concept, as well as the corresponding
techniques and design, of area-wise distributed sensing that can work for
both dynamic and static and macro and micro conditions. In this chapter,
we will introduce how to develop a comprehensive theoretical background
for analyzing and identifying structural parameters from point mode,
long-gauge, and area-wise distributed sensing data.
2  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.1  FRAMEWORK OF STRUCTURAL


IDENTIFICATION USING AREA-WISE DISTRIBUTED
SENSING DATA

In the previous chapter, we introduced the basic concept of area-wise dis-


tributed sensing of a structure, which was based on laying a long-gauge
strain sensor in the key, or critical, areas of the monitored body. In this
chapter, we will introduce a state identification method based on the
structure of area-wise distributed-strain data, combined with the material
properties, geometry, and other basic structural parameters, such as de-
flection, angular rotation, curvature, stiffness, internal stress, load, and so
on, to achieve the performance evaluation of the component level and the
complex structure. Currently, the standard method mainly includes two
levels of structure performance evaluation based on test data, especially
for bridges. The first level is regular testing for evaluating the index of
the structure and safety performance evaluation and then, depending on
the outcome of the first level of testing, evaluating the bearing capacity of
the bridge structure, such as the load carrying capacity evaluation test for
different load levels. This approach is only aimed at assessing the structural
load bearing capacity to evaluate the basic performance of the structure,
without taking the structural damage into consideration. It is clear that the
potential risk of damage propagation may lead to serious problems. Thus,
it is difficult to assess the structural performance in a meaningful way by
using this approach. By employing the method of area-wise distributed
strain in the critical or key areas, we are not using a single evaluation
index for measurement. Based on the area-wise distributed-strain data we
can obtain a variety of structural state parameters, which can be used to
analyze the structure through the entire data. We know that the structure
undergoes a long and slow process of degradation during its entire life
cycle. Taking a concrete structure as an example, in the early stage of
the structure, neutralization of concrete, which is the reaction of carbon
dioxide from the outside and cement hydration product such as calcium
hydroxide, is the main process, and the local stiffness degradation of the
structure is slow and not obvious. When concrete deteriorates to a certain
extent, steel gradually starts corroding and the stage of the development of
structural degradation begins. With the gradual deepening of the concrete
neutralization and steel corrosion, corrosion cracks appear, which accel-
erate the corrosion rate of the reinforcement. When the crack width is in-
creased, the steel bar is directly exposed to the air, so the deterioration rate
of the structure is further accelerated. This leads to the obvious damage
of the exposed steel bars and concrete. In the long-term monitoring pro-
cess, since the long-gauge strain sensor has high durability, it is suitable
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   3

for long-term strain monitoring. Thus, the utilization of the area-wise


distributed sensing system is an effective way to accurately analyze and
assess the material degradation process, and perform the qualitative and
quantitative analysis of the damage. The identification and localization of
micro damage, and the analysis of the support degradation are realized in
Section 1.6, based on strain modal analysis. The fiber model introduced
in Section 1.9 is based on the analysis of the deterioration of concrete and
the degree of corrosion of reinforcing steel bar. Following this approach,
we can further find the parameters of the structure in real time and accord-
ingly we can correct the structural analysis model, and subsequently we
can carry out internal analysis with a high degree of accuracy.
The overall flow chart that shows how the area-wise distributed sensing
data can be used to measure structural performance is shown in Figure 1.1.
This includes the following main steps. First, in the life cycle of structure,
deterioration of concrete, steel, and other materials is a slow process of
change, and then all parameters are obtained through the existing structure

Figure 1.1.  Framework of structural identification


4  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

parameters for the structure analysis of pre-damage. When the structure


appears abnormal, on the one hand, through the continuous monitoring
data in the long term, it is learned that the deterioration process from ma-
terial gradient to the mutation. On the other hand, the damage was found
by the distributed sensing areas and the damage location was determined.
In the quantitative analysis of damage, the deterioration of concrete and
the corrosion of steel bar are obtained by iterative calculation. Finally, the
modified parameters are incorporated in the finite element (FE) model to
perform the internal complex stress analysis, which is used to evaluate the
structural performance and the life span of the structure.

1.2  ROTATION AND DEFLECTION IDENTIFICATION

Deflection distribution is a critical indicator for the evaluation of struc-


tural performance. As one of the most important aspects of structural
health monitoring, deflection monitoring can accurately reflect the state
of internal forces in a structure, for instance, when a bridge is under both
static and dynamic load. Deflection is also an important parameter for the
indication of the accumulation of damage. In recent years, in particular,
with the growing number of long-span bridges being constructed globally,
deflection monitoring has become an increasingly indispensable method
for monitoring.
While numerous types of displacement sensors have been developed
for deflection measurement in long-span bridges, there are limitations in
their practical applications. For example, linear variable differential trans-
ducers, which are a type of contact sensor, can only measure deformation
accurately if a fixed base and a complex installation setup are available.
Another type of sensor that is used for deflection measurement is the non-
contact displacement transducer. Examples of this type of sensor include
the laser Doppler vibrometer, GPS sensors, vision-based approaches, and
ground-based interferometric radar technology (Fujino et al. 2000). These
sensors overcome the disadvantages of contact sensors, but their applica-
tion is still affected by several factors, including critical environmental
conditions, high equipment costs, and the complexity of data processing.
To address these challenging problems, various studies have been con-
ducted to obtain structural deflection information indirectly by using other
types of sensors. The double integration method (DIM) based on the use
of an accelerometer is commonly used to measure displacement (Smyth
and Wu 2007). Curvature and inclination-based approaches, which are
based on integration methods, have been applied to vertical displacement
measurements via the use of fiber-optic sensors. However, the integration
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   5

method is sensitive to the environmental and other sources of noise. In


addition, these sensing devices, when used for deflection calculations, are
“point” sensors and may not accurately reflect the local damage unless the
area where the sensor is located covers the damaged region exactly. With
the development of fiber-optic sensing technology (Hampshire and Adeli
2000; Tondini et  al. 2015), a long-gauge fiber-optic sensing technology
(Li and Wu 2007) has been developed to achieve the aim of distributed
sensing for structural health monitoring, which calculates the deflection
deformation using an improved conjugated beam method (CBM) (Shen
et  al. 2010). This deflection distribution identification method has been
used successfully in simple structures, including simply supported beams
and continuous beams under a pure bending moment.

1.2.1  INTEGRATION METHOD

Following the mechanics of materials, the structural deflection can be cal-


culated directly through the quadratic integral of the structural deflection.
In real-world projects of structural health monitoring, the monitoring data
usually includes noises, which adversely affect the accuracy of the com-
puted deflection, via quadruple numerical integration. Therefore, in order
to overcome this drawback, Yau and other experts developed a polynomial
function of curvature (strain) distribution that captures inherent physical
characteristics and features of the external loads. Therefore, the calcula-
tion of the deflection can be reduced to the evaluation of a quadratic inte-
gral of a polynomial function. The equation is as follows:

k ( x ) = cn x n + cn − 1 x n − 1 + ⋅ ⋅ ⋅ + c1 x + c0 . Eq. 1

In this equation, c0 , c1 , c2 ,..., cn are the parameters of curvature


function which can be gained by the regression analysis of the strain
­measurement; x is the curvilinear coordinates along the beam. Then, the
intersection angle or the deflection shape function can be calculated by
the  curvature function. The intersection angle is the first field integral
of the curvature and the deflection curve function is the quadratic integral
of the curvature. The equation is as follows:
cn n + 1 cn − 1 n c
θ ( x ) = ∫ k ( x )dx = x + x + ⋅ ⋅ ⋅ + 1 x 2 + c0 x + c′0 , Eq. 2
n+1 n 2
cn c
y ( x ) = ∫∫ k ( x )dx = xn+ 2 + n −1
x n +1
( n + 1) ∗ ( n + 2 ) n ∗ ( n + 1)
Eq. 3
c c
+ ⋅ ⋅ ⋅ + 1 x 3 + 0 x 2 + c′0 x + c″0 ,
2∗3 1∗ 2
6  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Where c0' and c0'' can be determined by incorporating the boundary


conditions.

1.2.2  CONJUGATED BEAM METHOD

The classical CBM was first presented by Otto Mohr in 1860. Essentially,
it requires the same amount of computation as the moment-area theorems
to determine a beam’s deformation. Compared with DIM, CBM can obtain
the structural deformation without measuring the slope of the boundary
points. An obvious disadvantage of classical CBM is its inapplicability in
cases with support settlement action or temperature variation. Derivation
includes an explicit linear function under simply supported conditions.
Since load is always unknown, or hard to measure precisely in actual struc-
tures, it is crucial to establish an equivalent simulation to replace an arbi-
trary load distribution.

1.2.2.1  Conjugated Beam Theory

According to mechanics of materials, a change or variation in the value


of the applied load results in a change in the bending moment distribution
along the axis of beam. Moreover, the absolute values of curvature distri-
bution at any point along the original beam are equal to the absolute val-
ues of uniform load distribution in the conjugated beam. Thus, the strain
distribution can be regarded as an appropriate simulation for any arbitrary
load condition in actual structures, as shown in Eq. 4:

k ( x ) = M ( x ) / EI = ε L ( x ) / y = q′ ( x ) , Eq. 4

where k(x), M(x), and e  L(x) are the curvature, moment, and strain distribu-
tion in the x-direction of the original beam, respectively. EI and y indicate
section stiffness and distance from sensor location to the inertial axis, re-
spectively. q′(x) is the uniform load distribution in the x-direction in the
conjugated beam, as shown in Figure 1.2.
The derivation process is based on the following assumptions. First,
the actual structures are simplified as Euler beams with small deflections.
Second, it is assumed that the stress–strain function of the structure’s ma-
terial is linear elastic, and the structural deformation is mainly caused by
the integration of small bending deflections of all sections. The main pur-
pose of these assumptions is to ensure the application of plane-section
assumption. The tensile strain and the upward deformation are defined
to be positive. A simply supported beam under investigation is shown in
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   7

Figure 1.2.  The simply supported beam and conjugated beam

Figure 1.2. The beam is not only the original beam, but also its conjugated
beam. It has a length L and uniform flexural rigidity EI. For the purpose
of this discussion the beam is artificially divided into n zones, denoted as
Element 1 to Element n, with a uniform length of l = L/n. The dotted line
in Figure 1.2 indicates the moment distribution under arbitrary load con-
ditions in the original beam, and the solid line denotes the equivalent load
distribution in the conjugated beam.
Eq. 4 can be transformed into:

kl = M l / EI i = ε lL / yi = M l / EI = q′i i = 1 ~ n, i=1 − n Eq. 5

where superscript “ ′ ” and “ – ” represent the parameters in the conjugated


beam and the average values of parameters in elements, respectively.
It is easy to obtain the left support reaction F ′ by calculating the bend-
ing moment at the right end of the conjugated beam:

l n ′  1
F′ = − ∑q
n i =1 i
⋅ n − i + 
 2

The moment M ′p of the boundary point between the pth element and
the p+1th element is:

p n ′  1 1 
M ′p = − l 2  ∑ i = 1 qi ⋅  n − i +  − ∑ i =1 qi ⋅  p − i + 2  
p ′
Eq. 6
n  2
( p = 1 + n ) .

The moment M ′p + 1/2 of the middle point of the ( p+1)th element is:

p n ′  1  1 1 
M ′p + 1/ 2 = − l 2  ∑ i = 1 qi ⋅  n − i +   p +  − ∑ i =1 qi ⋅  p − i + 2   .
p ′

n  2  2
Eq. 7
8  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

The displacement of the original beam is equal to the moment of the


conjugated beam in the same section:

υ Lp = M ′p , Eq. 8

υ Lp +1/ 2 = M ′p +1/ 2 . Eq. 9

Substituting (5) into (6) and (7), the following equations can be
obtained:
L2   p  n ε iL  1 p L
εi  1 
υ Lp = −
n2
   ∑  n − i +  −
2
∑ y  p − i + 2   , Eq. 10
 n i = 1 yi i =1 i 

L2  p + 1 / 2 n εi
L
 1 p L
εi 
υ Lp +1/ 2 = −    ∑  n − i +  − ∑ y ( p − i + 1) .
n2  n i = 1 yi 2 i =1 i 
Eq. 11

According to the CBM, the conjugated beam of a statically determi-


nate beam is also a statically determinate beam, but if the original beam
is a statically indeterminate beam, its conjugated beam will become an
“unstable mechanism” with a series of hinges. This variable mechanism
will not be stable unless under a special load condition that is equal to
the curvature distribution of the original indeterminate beam. Since the
same procedure is used for moment distribution in this variable mecha-
nism and in the corresponding statically determinate beam, the procedure
for monitoring the deformation of a statically indeterminate beam can be
simplified as an aggregate of monitoring the deformation of several sim-
ply supported beams. Otherwise, the accuracy of deformation monitoring
with CBM in one span of a continuous structure is related only to the
accuracy of strain measurements for the same span. It is free from accu-
mulation from strain measurement errors in any other span. Therefore,
compared with classical DIM, the remarkable characteristic of CBM can
readily restrict the inference of strain measurement errors to increase the
accuracy of deformation monitoring.

1.2.2.2  Extension with Combined Action from Loads


and Support Settlements

The purpose of this section is to provide a solution for the case which in-
cludes the effect of support settlements based on classical CBM.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   9

Figure 1.3.  Deformation distribution with support settlements. (a) One support
settlement, (b) two support settlements

Eq. 12 always holds:


ε L ( x)
V l ( x) = ∫∫ dxdy . Eq. 12
y
The simply supported beam under consideration is shown in
Figure 1.3a. It has a length L and settlement Δ in the right support. The
solid line and the dotted line denote the beam axis with and without sup-
port settlement, respectively. Due to no influence on strain distribution
from the settlement, e (x) = 0. Substituting this into Eq. 12, the structural
deflection is given as Eq. 13:

x
vL ( x) = ∆ . Eq. 13
l
This problem is again solved by classical CBM. Substituting e (x) = 0
into Eqs. 10 and 11, the formula mistakenly obtains that the deflection
along the beam is zero. Thus, it is concluded that classical CBM needs
improvement when support settlement is under consideration.
Considering combined action from loads and support settlements, the
deflection of the pth element can be divided into two parts. One is the
deflection from changes in strain distribution caused by loads and support
settlements, which can be calculated by Eqs. 10 and 11. The other is rigid
deflection from support settlements, which can be given as:
p
v Sp = ∆ r . Eq. 14
n
Assuming there are two settlements Δl, Δr in both supports in
Figure 1.3b, Eq. 14 can be modified as follows:
n− p p
v Sp = ∆ i + ∆ r , Eq. 15
n n
10  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.4.  Thermal expansion of a simply support beam

n− p −1/ 2 p +1/ 2
v Sp + 1/ 2 = ∆i + ∆ r . Eq. 16
n n
The deformation distribution can be obtained by Eqs. 17 and 18:

v p = v pL + v Sp , Eq. 17

v p + 1/ 2 = v pL+ 1/ 2 + v Sp + 1/ 2 . Eq. 18

The deformation from Eqs. 17 and 18 depends on only a few parame-


ters yi, n, p, L, Δl, Δr, which are the distance from the sensor location to the
inertial axis of the ith element, the total number of elements, the element
number of the point with unknown displacement, the length of the beam,
and the support settlements, respectively. All of these parameters are easy
to determine. As a result, both distributed fiber-optic sensors (FOSs) and
some displacement meter are needed for deformation monitoring. The
former is used to obtain the strain distribution, and the latter is used to
measure support settlements.

1.2.2.3  Applicability of CBM under Temperature Variation

In most of the previous studies, structures were usually assumed to be sub-


jected to constant temperature environmental conditions with load action,
but temperature variation must be taken into account in practical deforma-
tion monitoring.
Effect of temperature variation can be seen as a kind of equivalent load
action on the beam. Figure 1.4 shows the simply supported beam, which is
artificially divided into n zones, denoted as Element 1 to E ­ lement n. It is
assumed that on the upper surface and the bottom surface, the temperature
variations are Ti,u and Ti,b, respectively, for the ith element. According to
the Euler beam theory and can be transformed as follows:

εi ε iT,b − ε iT,u α
ki = = = Ti ,b − Ti ,u 
yi h h
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   11

where a, h, and eT(x) represent the coefficient of thermal expansion,


height of the section, and structural strain due to temperature variation,
respectively. Thus, the displacement of the boundary point between the pth
element and the p + 1th element can be given as:

α L2  p n  1  1
v Tp = − 2 
n h n
∑ i = 1
( Ti ,b − Ti ,u )  n − i +  − ∑ i = 1 (Ti ,b − Ti ,u )  p − i +   .
 2 
p
 2
Eq. 19

The displacement vTp + 1/2 of the middle point of the (  p + 1)th element is:

  1 
α L2  p + 1 / 2 n (Ti ,b − Ti ,u )  n − i + 2  .
v Tp + 1/ 2 =− 2
n h  n
∑ i =1 
− ∑ i = 1 (Ti ,b − Ti ,u ) ( p − i + 1)
p

Eq. 20

If there are combined actions from loads, support settlements, and


temperature variations, Eqs. 17 and 18 are modified as follows:

v p = v pL + v Sp + v Tp , Eq. 21

v p + 1/ 2 = v pL+ 1/ 2 + v Sp + 1/ 2 + v Tp + 1/ 2 , Eq. 22

where v Lp and v Lp + 1/2 can be calculated by Eqs. 10 and 11, v Sp and v Sp + 1/2
can be calculated by Eqs. 15 and 16, and v Tp and v Tp + 1/2 can be calculated
by Eqs. 19 and 20.

1.2.2.4  Extension under Continuous Beam Conditions

According to the CBM, the conjugated beam of a statically determinate


beam is also a statically determinate beam, but if the original beam is a
statically indeterminate beam, its conjugated beam will become an “unsta-
ble mechanism” with a series of hinges. This variable mechanism will not
be stable unless under a special load condition that is equal to the curvature
distribution of the original indeterminate beam. Because the same proce-
dure is used for moment distribution in this variable mechanism and in the
corresponding statically determinate beam, the procedure for monitoring
the deformation of a statically indeterminate beam can be simplified as
an aggregate of monitoring the deformation of several simply supported
beams. This means that Eqs. 17 and 18 are suitable for the deformation
monitoring of continuous beams. Because temperature variations can be
12  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

seen as some kind of equivalent load, Eqs. 21 and 22 are applicable for
the deformation monitoring of continuous beams with combined action of
loads and temperature variations.
Otherwise, the accuracy of deformation monitoring with CBM in one
span of a continuous structure is related only to the accuracy of strain
measurements for the same span. It is free from accumulation from strain
measurement errors in any other span. So compared with classical DIM,
the remarkable characteristic of CBM can readily restrict the inference
of strain measurement errors to increase in the accuracy of deformation
monitoring.

1.2.2.5  Experimental Verification of CBM

This experiment is designed to verify some theoretical conclusions men-


tioned in Part 4 that show, in theory, that CBM has more accuracy than
DIM under the same strain distribution with measurement errors. The ex-
periment also aims to demonstrate the effectiveness of Eqs. 17 and 18,
which are used to monitor structural deformation with combined action
from loads and support settlements.
The details of the continuous steel beam dimensions are illustrated
in Figure 1.8. A continuous steel beam with a total length of 2,400 mm is
uniformly divided into three spans, denoted as B1 to B3. Each span is then
uniformly divided into six elements. There are 18 elements in total, de-
noted as E1 to E18. The cross section has a rectangular shape. The width is
80 mm and the height is 8 mm. The Young’s module E and the moment of
inertia I are 210 GPa and 3.41 × 103 mm4, respectively. The displacements
observed from digital transducers (DTs) are regarded as the true values for
comparison with the monitoring data.
Two different loading modes (shown in Figure 1.9) are designed as
follows:

I. Concentrate loads F = 200 N, 400 N, 600 N are applied to the


middle point of B3.
II. Concentrate load F = 200 N is applied to the middle point of B2
with settlements of 25 mm in the two middle supports.

In each step, measurements are repeated five times to obtain the av-
erage strain.
According to Eqs. 10, 11, 15–18, the deflection formula can be ob-
tained by substituting L = 1,500 mm, y = 4 mm, and n = 6. For example,
the deflection formulas of P1, P2, and P3 in B1 are shown as Eqs. 23–25.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   13

And the deflection formulas of other monitoring points in B2 and B3 are


approximately the same as those of B1.

2 1
v1 = 2604 × ( 2ε1 + 6ε 2 + 7ε 3 + 5ε 4 + 3ε 5 + ε 6 ) + ∆ i + ∆ r , Eq. 23
3 3
1 1
v2 = 2604 × (1.5ε1 + 4.5ε 2 + 7.5ε 3 + 7.5ε 4 + 4.5ε 5 + 1.5ε 6 ) + ∆i + ∆r
2 2
Eq. 24
1 2
v3 = 2604 × ( ε1 + 3ε 2 + 5ε 3 + 7ε 4 + 6ε 5 + 2ε 6 ) + ∆ i + ∆ r . Eq. 25
3 3

For DIM, the initial rotations θ0 are −0.003, −0.0059, −0.009 when
F = 200 N, 400 N, 600 N in loading mode I, respectively. The q0 are
−0.0012 in loading mode II.
Figure 1.9 shows the differences in deflection in loading mode I from
DT measurement to various monitoring methods including Finite element
method (FEM), CBM, and DIM. In DIM, the monitoring deflection moves
away from the true deflection gradually, and the monitoring errors keep in-
creasing as more and more element strain data are used. In CBM, however,
the monitoring errors keep decreasing while the monitoring point moves
from P1 to P9. The maximum error is about only 5 percent, especially in
P7 to P9 of B3. The main reason for this is that strain measurement errors
in any span are restricted to have no influence on deflection of another
span in CBM, but in DIM, the accuracy of monitoring deflection relies on
all elements’ strain measurement data having enough accuracy. The strain
measurement errors from E1 to E6 produce uncontrolled influence in the
monitoring displacements of P4 to P9. This illustrates that DIM is unsuit-
able for structural deformation monitoring with strain measurement data
with considerable errors.
Figure 1.5 gives the difference in deflection between DT measure-
ment and two monitoring methods in loading mode II. The results from
CBM agree well with the true values with little errors. Thus, it is concluded
that Eqs. 17 and 18 can be used to monitor structural deformation with
the combined action of the loads and support settlements. Because of the
strain measurement error accumulation mentioned earlier, the monitoring
displacements from DIM have much difference from the true values, espe-
cially in P7 to P9. It can be concluded that CBM can provide more accur-
acy than DIM for practical structural deformation monitoring. However, in
view of the small strain variation in everyday monitoring, it is necessary to
reduce the strain measurement error of existing PPP-BOTDA technology.
14  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.5.  Deflection comparisons between different monitoring methods

1.3  STRAIN MODAL ANALYSIS THEORY

Although the mode theory based on acceleration measurement has a rel-


atively mature development, there are still a lot of limitations during the
application process of structural health monitoring. For example, the identi-
fication of the minute damages of the structure has been studied for 10 years
without any satisfactory results. Although the strain mode theory based on
strain measurement has developed to some extent, it is far less mature than
the acceleration mode analysis method. The reason is that traditional point
mode is sensitive to the structure’s local information, such as local pore and
grooving, so it is not appropriate for modal analysis with the purpose of
identifying the structural macro information. The output of the long-gauge
sensing technology introduced in Chapter 3 is the average strain within a
certain gauge length; therefore, it can reflect the structural macromodal in-
formation when reflecting structural local information. So the development
of long-gauge sensing technology and area-wise distributed sensing method
provides a good opportunity for the development of strain mode method.
Based on this opportunity, the authors’ team studied the strain mode theoret-
ical analysis based on long-gauge strain and have expanded the traditional
strain mode theory. Relative to traditional point mode strain mode theory,
the modal theory based on long-gauge strain can be called macro-strain
modal theory which is also called long-gauge strain modal theory. In this
book, it is called long-gauge strain modal theory in general.

1.3.1  POINT-TYPE STRAIN MODAL ANALYSIS

Before introducing long-gauge strain modal theory, we first introduce the


strain modal theory based on traditional point modal strain measurement.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   15

In physical reference frame, the equation of motion of a typical multi-­


degree-of-freedom (DOF) linear nontime-varying system is:

[ M ]{ x( t )} + [C ]{ x( t )} + [ k ]{ x ( t )} = { f ( t )} , Eq. 26

where [M], [C], and [K] are mass, damping and stiffness matrix, ­respectively.
{ f (t)} is the excitation vector. {ẍ(t)}, {ẋ(t)}, and {x(t)} are the displace-
ment, velocity, and acceleration response vectors of the structure.
After performing Laplace transform of both sides of the equation of
motion, Eq. 26 turns into:

([ M ] s 2 + [C ] s + [ k ]){ x ( s )} = { f ( s )} , Eq. 27

where  { F ( s )} and { X ( s )} are the Laplace transform of { f (t )} and { x ( t )} ,


respectively.
It can be concluded from (27) that:

{ x ( s )} = [ H d ( s )]{ f ( s )}, Eq. 28

where

[ H d ( s )] = ([ M ] s 2 + [C ] s 2 + [ K ])
−1
, Eq. 29

[ H d ( s )] is the displacement transfer function matrix of the structure.


According to structural dynamics, different damping systems have
different transfer modes of transfer function when the equations of motion
are transferred from the physical to modal coordinate systems. Viscous
damping model can be divided into a proportional damping system and
a general damping system. Structures with small damping can be consid-
ered as proportionally damped; otherwise, general damping system should
be chosen for modal analysis.
When a structure is defined as a proportionally damped system,
damping matrix can be expressed as follows:

[C ] = α [ M ] + β [ K ], Eq. 30

where α and β are proportional constants.


The orthogonality of the vibrating mode can lead to the following
equation:

diag ( M i ) = [ Φ ] [ M ][ Φ ] diag (Ci ) = [ Φ ] [ M ][ Φ ] diag ( Ki ) = [ Φ ] [ M ][ Φ ],


T T T

Eq. 31
16  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

where [ φ ] = {φ1}{φ2 }…{φN }  , and [ φ ] is the Mode Shape Matrix


 

that consists of all modal vectors {φi } (i = 1, 2, 3, . . ., N ). M i , Ci , and Ki


(i = 1, 2, 3, . . ., N ) are the modal mass, modal damping, and the modal
stiffness of order i, diag is the diagonal matrix. Therefore, the following
Generalized Mass, Damping and Stiffness matrices can be obtained:

[C ] = ([ Φ ]T )
−1
diag (Ci ) ([ Φ ]) ,
−1
Eq. 32

[ K ] = ([ Φ ]T )
−1
diag ( Ki ) ([ Φ ]) .
−1

Substituting (32) into (29) results in:


−1
[ H d ( s )] = ([ Φ ]T ) diag ( M i s 2 + Ci s + Ki ) ([ Φ ]) 
−1 −1


(
= [ Φ ] diag 1 / ( M i s 2 + Ci s + Ki ) [ Φ ]
T
) T

= ∑ i =1
{φi }{φi }T {φi }{φi }T
∑ i =1 M ( s 2 + 2ξ ω s + ω 2 ) ,
N N
=
M i s 2 + Ci s + Ki i i i i

Eq. 33
Ki Ci
where ω i = ,ξ = , and ω i  is the modal circular frequency,
Mi i 2 M iω i
or the natural frequency, of order i, and ξi is the modal damping ratio of
order i.
The transfer function matrix of acceleration when the system is static,
based on the Laplace transform, is:

s 2 {φi }{φi }
T

[ H a ( s )] = ∑ i = 1 M
N
. Eq. 34
i ( s 2 + 2ξiω i s + ω i2 )
By defining s = jω  ( j = −1), the matrix of complex frequency
­response function (FRF) of displacement and acceleration will be:

{φi }{φi }T
[ H d (ω )] = ∑ i =1 M
N
, Eq. 35
i (ω i + ω + 2 jξ iω iω )
2 2

−ω 2 {φi }{φi }
T

[ H a (ω )] = ∑ i =1 M
N
. Eq. 36
i (ω i2 − ω 2 + 2 jξiω iω )
Each element of the acceleration FRF matrix is:
−ω 2φ φ
∑ i =1 M (ω 2 − ω 2 pi+ 2qi jξ ω ω ) ,
N
H apq (ω ) = Eq. 37
i i i i
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   17

which is the parameter mode of viscous proportion damping system and


where p, q = 1, 2, . . ., n.
Substitute (32) into (28):

{φi }{φi }T
[ X ( s )] = ∑ i =1 M ( s 2 + 2ξ ω s + ω 2 ) [ F ( s )] .
N
Eq. 38
i i i i

For Euler beam,

1 ∂ 2 ω , Eq. 39
=
ρ ∂ x2
1
where is the curvature of the beam, ω is the deflection of the beam.
ρ
Since
z ∂ 2ω
εx = = z , Eq. 40
ρ ∂ x2
where ex is the strain of the beam along direction x, z is the distance from
the testing point to the neutral axis of the beam. Herein only positive strain
has been taken into account.
 N {φi }{φi }
T

∂ ∑ i =1
2
M i ( s 2 + 2ξiω i s + ω i2 )
[ F ( s ) ] 
∂ 2 [ X ( s )]  
[ ε x ( s )] =
∂x 2
z=
∂x 2
z

Eq. 41

where {φi } [ F ( s ) ] is the energy gained from the outside, the differential
T

is often treated as a constant. And the uniform beam can be acquired


through (41):

∂ 2 {φi }
z {φi }
T


[ε x ( s )] = ∑ i =1 M s 2 + 2ξ ω s + ω 2 [ F ( s )]
2
N x
Eq. 42
i( i i i )

∂ 2 {φi }
where {φi } is the modal displacement. Make z = {φiε } and {φiε }  
∂ x2
is the strain mode shape, so (42) can be written as:

{φiε }{φi }T
[ ε x ( s )] = ∑ i = 1 M [ F ( s )] .
N
Eq. 43
i ( s 2 + 2ξiω i s + ω i2 )
18  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Therefore, the strain transfer function matrix is:

{φiε }{φi }T
[ H ε ( s )] = ∑ i = 1 M
N
. Eq. 44
i ( s 2 + 2ξiω i s + ω i2 )
Make s =  jω , and the FRF matrix can be derived as follows:

{φiε }{φi }T
[ H ε (ω )] = ∑ i =1 M
N
. Eq. 45
i (ω i2 − ω 2 + 2 jξiω iω )

1.3.2  LONG-GAUGE STRAIN MODAL THEORY

Most vibration-based damage identification methods are based on the


assumption that the system can be treated as linear in the pre- and post-
damaged states, making the dynamic responses suitable for modal super-
position methods. Thus, herein, all analyses and discussions are based on
a linear time-invariant system.

1.3.2.1  Macro-Strain FRF

For a beam structure with two local DOFs, one for the vertical transla-
tional motion and the other for the rotation, at each node (Figure 1.6),
the macro-strain of several consecutive elements from the mth long-gauge
sensor with Lm gauge length can be obtained from the rotational DOF of
the first node of the first element (ith DOF) and that of the second node
of the last element (jth DOF) on a reasonable assumption that the distance
from the inertia axis to the sensor location of each element (demoted as
hm) is the same. This can be shown at a certain time and frequency as:
hm
ε m (t ) = ⋅ υi ( t ) − υ j ( t ) 
Lm 

hm
ε m (ω ) = ⋅ υi (ω ) − υ j (ω )  . Eq. 46
Lm 

Figure 1.6.  Beam structural model


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   19

Therefore, the macro-strain FRF between the measurement from the


mth sensor and the excitation at the pth DOF can be achieved by

ε m (ω )
ε
H mp (ω ) = . Eq. 47
Pp (ω )
where the bar “–” represents the macro-strain. Submitting Eq. 46 into Eq. 47
h
with ηm = m   can lead to
Lm
υi (ω ) − υ j (ω ) 
ε
H mp (ω ) = ηm ⋅ 
Pp (ω )
= ηm ⋅ H ipd (ω ) − H djp (ω ) , ( ) Eq. 48

where H ipd (ω ),  H djp (ω ) are displacement FRFs at the ith and jth DOF.
Eq. 48 can be further expressed as:

(
ηm ϕ ir − ϕ jr ϕ pr ) ηm ( Aipd − r Adjp )
∑ r =1 M (ω 2 − ω 2 + 2 jξ ω ω ) ∑ r =1 ω 2 − ω 2 + 2 jξ ω ω =∑ r =1 ω 2 −
N N r N
ε
H mp (ω ) = =
r r r r r r r r

ηm ( Aipd − r Adjp ) ε
Amp
∑ r =1 ω 2 − ω 2 + 2 jξ ω ω =∑ r =1
N r N r
= . Eq. 49
r r r ω r2 − ω + 2 jξ r ω r ω
2

where the macro-strain modal constant may be written as:

(
ηm ϕ ir − ϕ jr ϕ pr ) ϕ pr
r
ε
Amp = ηm ( r Aipd − r Adjp = ) Mr
=
Mr
δ mr , Eq. 50

with a given definition named by modal macro-strain (MMS) as

(
δ mr = ηm ϕ ir − ϕ jr . ) Eq. 51

The comparison of Eqs. (A-10) and (49) can lead to:

r H lpd (ω ) r Alpd ϕ lr ϕ lr
= = = , Eq. 52
r H mp (ω )
ε ε
r Amp δ mr (
ηm ϕ ir − ϕ jr )
where the subscript r is assigned to the rth mode. An important conclu-
sion can be drawn from this equation that the relation between displace-
ment and macro-strain FRFs is load and frequency independent but spatial
related.
Taking into account a single macro-strain FRF for the rth mode from
Eq. 49 it can be written as:

r
ε
H mp (ω ) = Rr H mp
ε
(ω ) + j ⋅ rI H mp
ε
(ω ) , Eq. 53
20  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

with the real and imaginary part expressed by:


 ω r2 − ω 2 
R ε
r H mp (ω ) = Amp
r   , Eq. 54
 (ω r2 − ω 2 ) + ( 2ξ r ω r ω ) 
2 2

 −2ξ r ω r ω 
I ε
r H mp (ω ) = Amp
r   . Eq. 55
 (ω r2 − ω 2 ) + ( 2ξ r ω r ω ) 
2 2

The magnitude and phase of the FRF may be shown as:

( (ω )) + ( rI H mp (ω )) =
r
2 2 Amp
ε
r H mp (ω ) = R ε
r H mp
ε
,
(ω r2 − ω ) + ( 2ξrω rω )2
2 2

Eq. 56

I ε
r H mp (ω )  −2ξ ω ω 
φ H = arctan = arctan  2 r r 2  , Eq. 57
R ε
r H mp (ω )  ωr − ω 

In contrast to displacement FRF, it can be seen that macro-strain FRF


differs in modal constant with respect to the real, imaginary part, and mag-
nitude representations and holds the same phase representation.

1.3.2.2  Modal Parameters Extraction

By comparing Eq. (A-10) with Eq. 49, it is realized that the expressions
of displacement and macro-strain FRFs share the same denominators and
only differ in numerators. It is well documented in numerous literature on
modal analysis (Ewins 2000) that resonant frequency and damping ratio
are uniquely determined by the denominators, whereas mode shapes are
determined by the numerators (i.e., modal constants) when using FRF. In
other words, with respect to the identification on resonant frequency and
damping ratio, displacement and macro-strain FRFs are equally effective.
It can be found that the displacement and macro-strain modal con-
stants have such relation as in Eq. 52, which assures that a similar param-
eter can be obtained from macro-strain measurements as mode shape from
displacements. For instance, concerning the direct identification method
based on the plot of magnitude macro-strain FRF in Eq. 56, the value at
each peak can be written as:
r
Amp ϕ pr
ε
r H mp (ω = ωr ) = = ⋅ δ mr . Eq. 58
2ξ r ω r 2 M r ξrω r
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   21

For a given mode, jpr /(2Mr xr wr) is a constant. Ignoring the amplitude


and only emphasizing the relative ratio of all components, the combina-
tion of macro-strain magnitude FRFs from all fiber Bragg grating (FBG)
sensors can lead to the construction of a vector as (59), and named as
modal macro-strain vector (MMSV).

{δ 1r , δ 2 r , ⋅ ⋅ ⋅, δ mr , ⋅ ⋅ ⋅}T Eq. 59

An important feature of MMSV is discussed herein. Let us suppose


M long-gauge sensors are installed onto the N-DOF beam in a distributed
manner (as in Figure 1.6). A macro-strain vector (MSV) of M-dimension
with each component from Eq. (46) can be assembled from displacement
vector (including vertical translation and rotation) by left-multiplying a
mapping matrix [B] as:



ε 1 ( t )   η1 ⋅ (υ i1 ( t ) − υ j1 ( t ))   υ1 ( t ) 
    
 ε 2 ( t )   η2 ⋅ (υ i 2 (t ) − υ j 2 (t ) ) 
  υ2 (t ) 
     
    
{ε ( t )} = 

= 

 = [ B ]M × N ⋅

 = [ B ]M × N ⋅ {υ l ( t )}
(υ ( t ) − υ jm ( t ))
m
 ε m ( t )   ηm ⋅ im   υl (t ) 
       
      
ε M (t )   η ⋅ υ N (t ) 

   M

(υ iM ( t ) − υ jM ( t )) 

 

Eq. 58

On the other hand, a complete set of MMS constants for the rth mode
with each component from Eq. 50 may be obtained as:

 r 

A1εp   η1 ⋅ ( r Aid1, p − r Adj1, p ) 


r A1dp 
 


ε
r A2 p
  η ⋅
  2
( d
r Ai 2, p − d
r A j 2, p ) 



d
r A2 p


  
{ r
ε
Amp } 
= 

ε
= 
 ηm ⋅ (

)

 =  B  M × N

⋅
 
d 
=  B∗  M × N ⋅ { r }
Alpd ,
, p − r A jm, p r Alp 
d d
 r mp 
A r Aim  
      
       
( )
d
 ε
r AMp
  ηM ⋅ r AiM , p − r A jM , p
d d   r ANp 
     

Eq. 59

For the convenience of statement, the subscripts [i1, i2, . . . iM] and
[j1, j2, . . . jM] for the representation of DOFs in Eqs. 58 and 59 are used
instead of the subscript l in vl  and ϕ lr without loss of the original physical
property. By comparing Eq. 58 with Eq. 59, we can find out that

[ B ]M × N =  B∗  M × N , Eq. 60
22  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Eq. 59 can then be expressed as:

{ r
ε
Amp}= [ B] ⋅ { r
d
ALp}. Eq. 61

Substituting Eqs. (A-11) and (50) into Eq. 61 will finally lead to
MMSV as:

{δ mr }M = [ B ]M × N ⋅ {ϕ lr }N , Eq. 62

Therefore, an important conclusion can be drawn from Eqs. 58, 61,


and 62 that the time-series and frequency responses of the measured
macro-strain versus displacement responses and the MMSV versus mode
shape share the same mapping relation [B].
The typical direct and logarithmic plots of the magnitude of a
macro-strain FRF obtained from the raw measurements are illustrated in
Figure 1.7. The acceleration magnitude FRFs under the same test are also
given for comparison. It is immediately clear from the direct plots that
the macro-strain FRF presents a more evident indication to the peaks at
low frequencies than acceleration FRF does. However, the amplitude of
macro-strain FRF decays heavily with frequency, the higher modes tend
not to show in the plots. This fact can be explained regarding the interrela-
tion of displacement and acceleration FRF:
ηm
  ε
  H mp (ω ) = ηm ⋅ ( H ipd (ω ) − H djp (ω )) =  ⋅ ( H ipa (ω ) − H ajp (ω )). Eq. 63
ω2
Obviously, in view of the graphical representation and modal iden-
tification, the macro-strain FRF is more sensitive to peak indication
than acceleration FRF when the resonant frequency is small as shown
in Figure 1.7. This property is very valuable for the online monitoring
of structures with high flexibility such as long-span bridges with initial
frequencies having very small quantity below 1 rad/s, even 0.1 rad/s and
densely congregating. From logarithmic plots it seems that by the direct

Figure 1.7.  Macro-strain and acceleration FRFs


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   23

Figure 1.8.  Long-gauge modal strain theory’s superiority over modal


displacement

use of the raw responses the measured macro-strain FRF is polluted by


heavier noise than acceleration FRF. Smoothing windows should be care-
fully selected for data processing.
As mentioned earlier, the long-gauge modal strain theory can obtain
accurate global modal properties such as frequencies, mode shapes, damp-
ing ratios, and phases, which fulfill the same functions of modal accel-
eration theory. However, modal acceleration theory has some limitations
in identifying long-gauge strain modal information. For instance, long-
gauge modal strain shape cannot be precisely calculated by displacement
mode shape while displacement mode shape can be successfully obtained
by long-gauge strain mode shape under 10 percent noise as shown in
Figure 1.8.
Theory: (a) long-gauge strain mode shape with 10 percent noise is
able to derive displacement mode shape; (b) displacement mode shape is
unable to derive long-gauge strain mode shape.

1.4  NEUTRAL AXIS HEIGHT IDENTIFICATION

The neutral axis depth (NAD) of a beam member is an important parame-


ter for structural design and structural safety evaluation. A few approaches
have been developed in the literature for NAD estimation using static or
24  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

dynamic strain measurements. In this section, a new method to determine


the neutral axis position is developed by using a concept of modal strain.
In the proposed method, macro-strain time histories are first measured
by applying long-gauge FBG sensors, then the modal strain is generated
from the measured dynamic macro-strain through Fourier Transform (FT).
Finally, the NAD is determined from the produced modal strain. The NAD
and the neutral axis positional coefficient are found to be very sensitive
to structural damage; thus, they are appropriate indicators for structural
damage detection.

1.4.1  NEW METHOD TO DETERMINE THE NEUTRAL AXIS


LOCATION

For beam-like structures, the assumption of plane section, as shown in


Figure 1.9, is common. It means that as long as the maximum bending
stress is less than the yield stress, any deformation due to shear across the
cross section is ignored and the strain will change linearly. Based on this
assumption, the neutral axis positional coefficient, y, is applied to express
this linear relationship:

ε ( t ) = ψε ′ ( t ) , Eq. 64

where e(t) and e′(t) are the strains at different depths of the same section. With
basic knowledge of geometry, the depth of the neutral axis h can be easily ob-
tained since the strain is zero at the location where the neutral axis lies. There
are two cases for the strain measurement: (1) sensors installed at different sides
of the neutral axis; (2) sensors installed at the same side of the neutral axis.
For these two cases, the equations to calculate the depth of the neutral axis are
shown as Eqs. 65 and 66, respectively, where H means the vertical distance
between the two locations where sensors are installed.

Figure 1.9.  Strain distribution along the cross section with sensors installed at:
(a) different sides of the neutral axis; (b) the same side of the neutral axis
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   25

ψ
h= H , Eq. 65
ψ +1

ψ
h= H . Eq. 66
ψ −1

Unlike using the maximum or averaged strain values of the measured


strain time history in the literature, the modal analysis of the strain time
history is first performed in the proposed method to determine the loca-
tion of the neural axis. As well known, the data in the time domain can be
transformed into those in the frequency domain with the FT. Therefore, the
relationship between the strain e(t) and e′(t) in the frequency zone can be
obtained through Eq. 67:
+∞ +∞ +∞
ε (ω ) = ∫−∞ ε (t )e − iω t dt = ∫−∞ ψε ′(t )e − iω t dt = ψ ∫−∞ ε ′(t )e − iω t dt = ψε ′(ω ) ,
Eq. 67

where e(w) and e′(w), named modal strains in the frequency domain, are
the counterparts of e(t) and e′(t) in the time domain. The peak value of
e(w) or e′(w) corresponding to the structure’s natural frequency is applied
to calculate the coefficient y as shown in Eq. 68, where e(wr) and e′(wr) are
the modal strain of the rth order mode. The 1st bending mode is a­ pplied
throughout this chapter as it can be easily and accurately identified:

ε ( wr ) , Eq. 68
ψ =
ε ′( wr )
The coefficient y contains enough spatial information of the neutral
axis from which the neutral axis position is estimated. If damage occurs
within the monitored zone, the neutral axis will move and the coefficient
y will change as well. Therefore, this coefficient has the potential to be a
new index for structural damage detection.

1.4.2  THE PROCEDURE TO DETERMINE THE NEUTRAL


AXIS LOCATION VIA MODAL STRAIN

For the readers’ easy understanding, the framework of the proposed method
is illustrated in Figure 1.10. It includes four main steps: (1) acquiring dy-
namic strain data with the installed strain sensing and data acquisition
system; (2) obtaining the peak value at the selected natural frequency of
26  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.10.  Procedure to determine neutral axis location with modal strain and
implement an SHM scheme

the frequency spectrum, which is calculated from the strain time history
through Fast Fourier Transform (FFT); (3) extracting the position coeffi-
cient of neutral axis y through a linear fitting between the modal strains of
two sensors within the same cross section, namely the slope of the fitting
line, and (4) extrapolating the neutral axis position with Eqs. 65 or 66 us-
ing the spatial information of sensor installation. There are also other ways
to determine the neutral axis location. As shown in Figure 1.10, from the
modal strain distribution along the cross section, the neutral axis location
can be determined by finding the line that is perpendicular to the vertical
axis of the cross section and passes through the zero value of the modal
strain. This is an equally effective method as the methods described in
the above steps. After the coefficient y and the neural axis position are
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   27

estimated, both of them can serve as damage detection indexes since they
are sensitive to structural damage.

1.4.3  EXPERIMENTAL EXAMPLE

Neutral axis positions of the girders of a bridge are studied by record-


ing macro-strain time histories. The investigated bridge in northern New
­Jersey of the United States is a multigirder steel stringer bridge constructed
in 1984. It is comprised of four spans using a standard steel stringer design
of girders (Figure 1.11).
The investigated bridge has four spans and we focused on the second
span on the southbound side for this field test due to time and cost limit-
ations. This span has the skew on one side. It passes through a park and
allows for unrestricted underside access. To monitor this span by using the
developed LG-FBG area-wise sensors, the gauge length of the sensor is
designed to be 1 m, and a total of 44 sensors are deployed on two girders
of this span as shown in Figure 1.12. The first 10 sensors are mounted
on the 25 and 50 cm along the web of the center area of the third girder
(Figure 1.13). They will be used for girder neutral axis determination,
which will be presented later. Another 34 sensors are mounted on the top
surface of the bottom flanges of girders 3 and 6 to measure their distribut-
ed-strain responses under ambient excitation.

Figure 1.11.  Profile of a steel-concrete composite beam bridge


28  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

9.91m 9.91m 9.91m


1 2 3 4 5

10 Sensors LG-FBG
(SE U.)
4 Sensors 9 Sensors 4 Sensors Girder 3
17.08m Strain gauge

4 Sensors 9 Sensors 4 Sensors Girder 6

1.5m F6-F10
F5-F1
F24-F27 F15-F23 F11-F14
37.64m
Girder 3

1.5m
F24-F27 F15-F23 F11-F14
33.15m
Girder 6

Figure 1.12.  Instrument plan

Figure 1.13.  Typical sensor layout on a section


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   29

All the LG-FBG sensors are manufactured in the lab including o­ ptic
fiber prestressing, packaging, calibrating, and heat sealing. Field test
­included determining the gauge location precisely, removing the paint and
corrosion, gluing the gauge in place with epoxy, and running the cable
along the girder to the designated data acquisition system. The epoxy
becomes firm within a few minutes, making these gauges effectively a
permanent installation. The bond is typically very strong and the gauge
can be used with confidence over a long time. This facilitates long-term
monitoring or future testing. The SM-130 is used as the data acquisition
system. The sampling rate for dynamic data collection is set to be 500 Hz.
During the ambient vibration test of the benchmark bridge, 30 and 60
date sets were recorded by the LG-FBG area sensors on June 8–9, 2011,
respectively. Each data set has the length around 10 minutes. To verify the
accuracy of the developed sensors, the recorded strain time histories are
compared with those from traditional strain gauges. The traditional strain
gauge used is the Hi-Tec weldable quarter-bridge strain gauge, which has a
2 inch shim length and a 1 inch gauge length (Weidner et al. 2011).
Figure 1.13 shows the typical longitudinal configuration of some
Hi-Tec strain gauges that coincide with the LG-FBG sensors on girders
3 and 6. These two kinds of sensors were set to collect data simultaneously
during the ambient vibration test in order to compare the results from these
two devices. The macro-strain time history at the line 2 of girder 3 re-
corded by the LG-FBG sensor 24 (Figure 1.12) is plotted in Figure 1.14a,
where the strain time history from the traditional gauge at the same time

Figure 1.14.  Strain time history comparison: (a) line 2 of girder 3, (b) line 2 of
girder 6
30  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

and location is also plotted for comparison. It is seen that peaks of two time
histories have very close magnitudes. To see the details, a time window
from 980 seconds to 1030 seconds is plotted in Figure 1.14a. It is clear that
the free-vibration responses excited by the traffic are successfully recorded
by both kinds of sensors and they are almost the same. The strain time
histories plotted in the figure are preprocessed data with a wavelet filter
(need to explain what type of wavelet) for observation noise reduction.
Similarly, Figure 1.14b shows the dynamic strains from the LG-FBG sen-
sor and the Hi-Tec strain gauge mounted at line 2 of girder 6, which further
illustrates that the developed sensors have the same capacity for dynamic
strain measurement as the widely applied traditional strain gauges. When
the structural elements that are measured are in good condition, for in-
stance the benchmark bridge studied in this case, the LG-FBG sensor will
output results similar to those of common sensors. It is expected, however,
that output from the LG-FBG sensors will be different from the traditional
strain gauge when structural damage exists. This is due to the fact that the
LG-FBG sensor is able to detect cracks within the long-gauge length (1 m
in this study), while the traditional gauge can only measure the local strain
(within 1 m in length). Especially, the newly developed sensors can also
be connected to make a sensor array to carry out distributed monitoring of
the entire structure at least in important areas, as performed in this study.
The neutral axis position is also investigated. Instrument plan in
­Figure 1.13 shows that two traditional strain gauges were mounted at the
top of the bottom flange and 50.4 cm up the web in a typical strain gauge
layout configuration (Weidner et al. 2011). Three static load cases by us-
ing six trucks with full loads, three trucks with full loads, and three empty
trucks were considered, and the loading locations are 1/4, 1/2, and 3/4
positions of the southbound span 2, respectively. The strain profile of a

Figure 1.15.  Height of neutral axis


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   31

Table 1.1  Neutral axis position estimation

Averaged height Standard deviation


Ambient test
123.9 4.62
Static test 3/4 span 120.1 —
with the
1/2 span 122.0 —
load at
1/4 span 126.2 —

given section is plotted by employing the static strain measurements. For


instance, Figure 1.15 shows the predicted strain profile of the cross section
at line 2 of girder 3 from the static truck test. It is seen that when the static
force location is fixed, the neutral axis position estimated from the strain
profile is stable even when the static load is different. However, the esti-
mated neutral axis position varies when the static truck load changes loca-
tions. This is probably because the heavy trucks at different locations cause
the deck and girders to have different connection conditions. Namely, the
property of the deck–girders has been changed with the truck locations
varying (Weidner et al. 2011). The estimated neutral axis position of the
cross section at the center of girder 3 from both the static truck tests and
dynamic macro-strain measurements is summarized in Table 1.1. It is seen
that the results from two methods are comparable, but it is obvious that
the method using the dynamic macro-strain time histories does not require
bridge closure during the test and it is much more convenient.
A method for estimating the neutral axis position from dynamic strain
measurement is proposed based on the modal analysis. To implement it,
modal analysis is first performed to transform the measured macro-strain
time history acquired with the proposed long-gauge macro-strain sensors
to the frequency spectrum through FFT. Subsequently, the peak values
of the frequency spectrum, namely the modal strain, are selected at the
first-order mode. Then the positional coefficient of neutral axis is obtained,
with which the NAD is extrapolated. FE simulation and experiments have
been performed to verify the proposed method. Both the simulation and
experimental examples have illustrated that the proposed method is pow-
erful for determining the neural axis position. Compared with the tradi-
tional methods for determining the neutral axis position with the strain
time history, the proposed method based on strain modal analysis in this
chapter presents a superior stability, especially, when some data acquisi-
tion noise is considered in the simulation. This has also been verified by
the correlation coefficient for the linear fitting being larger than 0.999 in
both the simulations and experiments.
32  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.5  STRUCTURAL DAMAGE IDENTIFICATION

Health and safety of major civil engineering structures such as long-span


bridges, high-rise buildings, hydraulic power plants, tunnels, and so on
play a critical role in safeguarding people’s lives, protecting societal as-
sets, and assuring a healthy and prosperous economy. These infrastructures
in service will inevitably suffer from environmental deterioration, traffic
overload, natural disasters, the aging of materials, and the degradation of
the overall performance caused by long-term usage and dynamic loads.
For instance, the most common structural damage in concrete structures
mainly includes the following types: first, crack problems in concrete; sec-
ond, corrosion of reinforcing steel bar; third, support or foundation settle-
ment problems; fourth, prestress loss of drag line or pretension cables. The
structural degradation caused by the above problems causes the reduc-
tion and deterioration in the structure’s ability to resist natural disasters,
normal loads, and other environmental impacts. Once the damage of the
key structural components begins to accumulate to a certain extent, if the
damage cannot be detected and retrofitted in time, the damage will rap-
idly expand, resulting in the destruction of the entire structure. Therefore,
in order to ensure the safety, integrity, and durability of the structure, it
is important, and urgently needed, to monitor and evaluate the degree of
damage and safety of the existing civil engineering structures that have
been in use for several years.
Damage identification of structures refers to any change of the struc-
ture’s characteristics. Therefore, if the applied load and the structure re-
sponse are known, structural damage identification can be carried out.
But the problem is that the applied load or its exact nature or value is
generally unknown. In view of this problem, structural damage detec-
tion in structural health monitoring has been done only based on struc-
tural response. For example, when conducting the vibration analysis of
a structure under environmental impacts, the exact nature or the input
force due to environment is unknown. However, by assuming that the en-
vironmental input load is a white noise, modal analysis can be conducted
to obtain the modal parameters of the structure, and structural damage
identification can be carried out. In this section, the method of dam-
age ­identification for structures under unknown loads is presented. The
­real-time monitoring and evaluation of structural degradation can be car-
ried out by the strain measurement in area-wise distribution. For concrete
structures, the damage measures or indicators can be divided into two
categories. The first type of damage are those that can be detected by the
direct use of monitoring data for the degradation assessment of structural
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   33

performance, such as structural crack detection and width identification


(Zhang and Wu 2012). The second type of damage are more complex
damage types such as loss of bond between the rebar and concrete due to
slippage (shear pinching), corrosion, etc., that can be analyzed based on
the monitoring data and the assessment of the degradation type or perfor-
mance (Hong et al. 2010; Hong et al. 2012).

1.5.1  CRACK WIDTH MONITORING

Health monitoring for reinforced concrete (RC) structures often includes


two main concerns on local damage detection and global behavior of the
structure. Due to low tensile strength of the material, the degradation
process of RC structures is always accompanied by the occurrence and
propagation of cracks. Moreover, a crack or crack pattern is an indicator
that shows the presence or the potential development of other degradation
such as concrete fatigue, alkali silica reaction, freeze–thaw, steel rein-
forcement corrosion associated with the leakage of water and chloride
through cracks. Therefore, the local damage detection including the mon-
itoring on the presence, location, and significance of cracks, reinforce-
ment corrosion, material behaviors or local stress/strain concentrations,
and so on are of engineering importance in maintenance and retrofitting
of concrete structures in service, or in particular after the occurrence of
natural hazards. On the other hand, global behavior evaluation provides
the valuable macroscopic information on the full structural assessment
and diagnosis.
This section deals with an application of the proposed sensors for
long-gauge area-wise distributed monitoring in structural assessment of
RC beams. We will also propose an integrated SHM strategy for RC flex-
ural structures, in this subsection, based on the classic bending theory and
Euler beam planar bending assumption.
For concrete structures, the monitoring of cracks is very important,
but the traditional resistance meter, based on crack meter or ductility meter,
is difficult to be used for continuous monitoring or a timely detection of
cracks. Based on the developed long-gauge fiber sensing technology, the
authors developed a crack detection method and a corresponding method
of width identification utilizing the measured long-gauge strain. This
approach has the following characteristics: (i) online monitoring; (ii) mon-
itoring is highly automated and does not require excessive human involve-
ment; (iii) monitoring is well distributed, and an optical fiber can monitor
the presence and development of cracks in the entire concrete structure.
34  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

When the sensor is fully affixed, the crack width can be calculated
by using the bond-slip theory, and the strain change of the concrete is
neglected in the calculation. The formula is as follows:

∫ ε ∗ dl n
ω ≈ εsl = l
l
*l = ∑ 1 / 2 * (ε i + ε i +1 ) * ∆l , Eq. 69
i =1

where ω is the width of the crack,  l is the length of the affected area, ε i is
the strain at the ith sampling point in the affected area, and Δl is the spatial
spacing of the samples.
When we adopt the width of the crack ω calculation, we use the fol-
lowing formula:

ω ≈ ε * L , Eq. 70

where ε is the average strain, and L is the length of the long-gauge sensor.
The average crack width within a certain region can be regarded as the
product of the increase of measured macro-strains and the gauge length of
the sensor under the constant load. The changes of measured data and the
corresponding average crack widths obtained by crack gauges and FBG
sensors of 200 mm gauge length are shown in Figure 1.16. The results
from these two different sensors present a good agreement. C1 represents
the crack width measured by the crack meter, and F1 represents the crack
width measured by the long-gauge sensor.

1.5.2  CORROSION MONITORING

Corrosion of steel in RC structures is a major problem worldwide, par-


ticularly in structures exposed to corrosive environments. A reliable

Figure 1.16.  Average crack results


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   35

inspection method for corrosion analysis is required at an early stage


before the functionality of the RC structure is seriously damaged (Zaki
et  al. 2015). Monitoring corrosion based on strain sensing techniques
has been widely observed in recent years (Geng, Wu, and Zhao 2009;
Lee, Yun, and Yoon 2009; Zhao et al. 2011; Zheng, Lei, and Sun 2001;
Zheng, Sun, and Lei 2009). In these studies, the main parameter used to
indicate the level of corrosion is the measurement of the expansion in
bar diameter due to corrosion deposits by winding FBG sensors around
the polished steel rebar in corroded areas. The following are the lim-
itations of using the commonly used FBG-based methods in corrosion
monitoring: (a) too local to use for large-scale structures; (b) vulnerable
when corrosion accumulates; (c) the corrosion can affect the fixation
point; (d) a temperature compensation sensor is required for accurate
measurements.
Corrosion of main steel reinforcements is one of the foremost causes
of early deterioration and shortening of service life of RC structures. Early
detection of corrosion can aid in reducing the maintenance cost and ex-
tending the service life of structures. The use of long-gauge strain sen-
sors in the field of corrosion monitoring of steel reinforcements in RC
structures offers the following advantages: (1) As mentioned in previous
chapters, the long-gauge sensor is a long-gauge strain measurement that is
suitable for macro-strain measurements of large-scale structures; this is in
contrast to short-gauge sensors, which are too local to detect the corrosion
damages. (2) The long-gauge sensor is a durable and rustproof material
that is unaffected by the corrosion of the main steel reinforcement. (3) The
packaged long-gauge sensor is a high-strength material that can function
for a long time without deterioration. This chapter proposes an approach
for early corrosion monitoring of steel reinforcements in concrete struc-
tures by using strain measurements of long-gauge packaged fiber Bragg
grating (PFBG) sensors.
Corrosion of steel bars that are embedded in concrete is the prin-
cipal cause of deterioration of RC structures. It reduces the structural
integrity of concrete by the loss in the area of steel, cracking of the cover
concrete, as well as the loss in the bond between the corroding steel and
the surrounding concrete. In this section, three different approaches,
continuous-strain ratio (CSR), distributed-strain ratio (DSR), and sec-
tion fiber model (SFM), were introduced to evaluate corrosion levels
using strain long-gauge measurements. The different groups of distrib-
uted-strain measurement sensors on the tension and compression parts
of concrete surface and steel reinforcements of RC beam were used to
determine different corrosion levels using the three different approaches
in the cases under consideration.
36  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.5.2.1  Continuous-Strain Ratio-Based Corrosion Identification


Technique

Corrosion in steel mostly ensues after the occurrence of cracks; there-


after, as the corrosion progresses, the stiffness of the beam is affected,
especially around the corroded region. In the case of localized corrosion,
all distributed strains along the steel reinforcements will increase with the
advancement of corrosion resulting from increments in the beam’s curva-
ture, according to values based on the percentage of damages in the cross
section of the beam. For a certain cross section of the beam, the measured
strains in the intact, cracked, and corroded states can be expressed as

M ∗ y M ∗ ycr M ∗ yco
ε = , ε cr = , ε co = , Eq. 71
EI  EI cr EI co

where M is the applied moment, and ε ,  EI ,  and y are the flexural strain,
flexural stiffness, and height of the neutral axis at a certain position, re-
spectively. The subscripts (o), (cr), and (co) represent the intact, cracked,
and corroded states, respectively. The variation in the position of the neu-
tral axis after initiation of crack, with low levels of progress, is small.
Therefore, we can assume that   ycr ≅   yco . Under a constant applied mo-
ment, the corrosion levels in a certain location based on the CSR approach
at a certain time can be expressed as
EI cr − EI co ε − ε cr
ψ = ∗ 100 = co ∗ 100, Eq. 72
EI cr ε co
where y represents the corrosion level in the cross section of the
beam. This ratio is probably affected by the variations in the curvature that
occur in a certain location with advancing localized corrosion.

1.5.2.2  Distributed-Strain Ratio-Based Corrosion Identification


Technique

To eliminate the effect of curvature variations, a DSR-based corrosion


identification technique is suggested by comparing the strains of two dif-
ferent sensors at two different locations: one at the target location and
another at the location of a reference sensor. The reference sensor will be
placed at a location with a low probability of crack and corrosion occur-
rence, which is usually near the support, as shown in Figure 1.17a.
This study considers the fact that the effective corrosion of the steel
bars inside the RC beams starts after the initiation of cracks under service
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   37

Figure 1.17.  Fundamentals of DSR approach: (a) configuration of sensors and


reference sensor, (b) relationship between reference and target strains

loads, before which there are no significant changes in the steel strains.
The following can be concluded from Figure 1.17b:

a. For the intact state, the strain at any section solely depends on the
magnitude of the moment ( EI i = EI r ), where the ratio between the
reference strain (er) and the target strain (ei) can be expressed using
Eq. 71 as
Mi ⋅ y
 εi  EI i Mi
 ε  = M r ⋅ y = a1 where a1 =
Mr
, Eq. 73
r 
EI r

where EI i and EI r are the flexural stiffness of the cross section at the tar-
get and reference locations, respectively, and the subscript (o) represents
the intact case. The slope of this state (a1) represents the ratio of the
moment.
38  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

b. For the cracked state, the strain at any section depends on the
moment ratio and change in the flexural stiffness ( EI icr = βicr EI r );
herein, the regression line shifts to a new position with an increase
in the slope of a2.
Mi ⋅ y
 εi  EI icr  1 
 ε  = M r ⋅ y = a2 where a2 = a1  cr  .
 βi 
Eq. 74
r cr
EI r

The subscript cr represents the cracked state, and βicr is the damage
index, which represents the ratio between the flexural stiffness of the
­uncracked and cracked conditions where 0hβicr ≤ r .
c. For the corroded state, after crack occurrence, we assume that the
ratio between the target and reference strains remains constant
under the same load, while, if corrosion starts at a certain location,
additionally the stiffness will be reduced gradually. ( EI ico = βico EI r ).
The regression line shifts with an increase in the slope of a3 are as
follows:
Mi ⋅ y
 εi  EI ico  1 
 ε  = M r ⋅ y = a3 where a3 = a1  co  , Eq. 75
r co  βi 
EI r
where subscript co represents the corroded state. Ultimately, the corrosion
level y defined in Eq. (72) can be determined.
In case corrosion is absent, the slope will not change, where a3 = a2
and ψ = 0.

1.5.2.3  SFM-Based Corrosion Identification Technique

To calculate the actual decrease in cross-sectional area of steel reinforce-


ment from measured strains, nonlinear analysis based on SFM method is
carried out regarding the class bending theory using the measured tensile
strains of steel and compressive strains of the upper surface of concrete
beam. Based on the assumption that plane sections remain plane and nor-
mal to the longitudinal axis, the nonlinear behavior of RC beam disre-
garding shear and bond-slip can derive from the constitutive relations of
concrete and reinforcing steel fibers into which each section is divided as
follows:
n
∑ N = ∑ σ i Ai + σ s As + σ s′ As′
i =1
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   39

In sum, the steps of SFM-based corrosion identification technique


to calculate the actual reduction in area steel with corrosion progress are:

a. Calculate the strain for every fiber over the whole section using the
strains of both tensile steel bar and compression fibers of concrete.
b. Calculate the stresses for every fiber over the whole section.
c. Take ΣN = 0 for equilibrium. Calculate as with the progress of
corrosion.
d. Finally, the corrosion damage y represented by the decrease in area
of steel is as follows:
 A − Asc 
ψ =  s0 * 100
 As 0 

1.5.2.4  Corrosion Monitoring of RC Structures

In this section, an experimental study is present to explain how to use dis-


tributed long-gauge strain sensors to build a corrosion monitoring system.
Two main cases will be introduced. The first case, when the corrosion
occurs inside concrete beam before cracks under low level of sustained
load (about 10 percent of the ultimate design load). The second case when
the corrosion occurs with slight cracks on the concrete under a sustained
load of about 30 percent of the ultimate design load. The two beams under
study with the same steel reinforcement and the same size with the same
levels of corrosion. Three different approaches, CSR, DSR, and SFM,
were proposed to evaluate corrosion levels using strain measurements.
The different groups of distributed PFBG sensors were installed on the
tension fibers of concrete surface and steel reinforcements of RC beam.
The distributed measured strains utilized to evaluate the deterioration of
the structures health with the advance of corrosion time. The corrosion
degrees calculated using each approach was compared with the weight
loss of external model. Finally, the total comparison between the two cases
under considerations and the ability of each approach to calculate the cor-
rosion levels in each case.
The PFBG sensors were prepared as mentioned before with a gauge
length of 250 mm, thereafter they fixed from its all length using the epoxy
resin on the designed locations and divided into four groups as follows:

1. SC group: this group consists of six sets of PCFL and another of


FBG sensors which were mounted on the corroded tension bar.
40  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

2. S group: this group consists of only six sets of PCFL sensors


mounted on the other un-corroded tension bar.
3. CB group: in this group six sets of PCFL and PFBG sensors were
fixed on the bottom surface of the RC beam.
4. CT group: only four sets of PCFL and PFBG sensors were installed
on the compression zone of the RC beam at a height of 160 mm
from the bottom surface.

Figure 1.18 shows the load and measurement of displacements before


and during the corrosion process of one steel tension bar of the RC beam.
It is clearly found that, the values of deflections were hardly increasing
with the advance of corrosion because at this stage the effect of steel bars
is not significant in the overall behavior of the beam at this stage At the
early stage of loading before cracks the behavior of the beam depends
mainly on the concrete properties till the stage of cracks initiation. As
well, there are no cracks caused by corrosion of the steel bar during the
corrosion test.
Figure 1.19(a) and (b) present the measured strains of long-gauge
PCFL sensors distribution of, the corroded bar (SC group), and the
un-corroded bar (S group), respectively. As it is clear from these figures,
all the long-gauge strain measurements along the two steel bars have very
small changes in strains during the corrosion process. In contrast, the
strains in the corrosion location have larger increasing trend than the oth-
ers resulting from the reduction in area steel in this location. Although the
changes in strains are very small, it can represent corrosion values because

Figure 1.18.  Load and displacement during corrosion process


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   41

the initial strains before corrosion are very small too. In that case we can
get the accurate values of strains in each loading stage because the applied
load is less than the crack load that mean the concrete still in the elastic
stage only the decreasing in diameter affect the strain values. The recorded
distributed-strain values of the bottom surface of concrete (CB group) are
shown in Figure 1.19c. The same trend as the measured strains of steel bar
was reviled; however the strain values before the onset of corrosion were
higher than those corresponding of steel bars, owing to the differences in
the location of strain measured from the neutral axis of the beam. Finally
the compressive strain distribution of the compressive fibers of concrete
surface (CT group) is shown in Figure 1.19d. The same trend was found in
the compression zone also the strain values at the corrosion section have
the largest increasing trend with the progress of corrosion.
The corrosion levels values were calculated using CSR approach for
each group. Figure 1.20a–c give the calculated corrosion levels along the
beam by using the measured distributed strains of (SC), (S), and (CB),
­respectively. From these figures it can be noted that, all the cross sections

Figure 1.19.  Monitoring results of (a) corroded bar, (b) un-corroded bar,
(c) tension concrete surface, and (d) compression concrete surface
42  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.20.  Corrosion levels along the length of the beam based on CSR for
(a) SC, (b) S, and (c) CB

of the beam gave a corrosion indications with different values in which the
maximum value was found at corrosion location. These results confirm
that, the calculated corrosion values affected by the curvature variations
as mentioned before. This method is not reliable to localize the actual lo-
cations of corrosion in this stage.
Second, the corrosion levels were calculated with DSR approach from
the distributed strains for SC, S, and CB sensors separately and compared
with each other as clarified in Figure 1.21a–d respectively. It can clearly
observe that, (1) the corrosion damages locations can be found obviously,
(2) by using this method the effect of curvature variations can be elimin-
ated and it can be considered that the calculated values related mainly to
the decrease in area steel, and (3) the corrosion levels calculated from the
strain measurements of both steel bars have convergent values in contrast
the corresponding values of the concrete surface have divergent values.
Finally, based on the SFM approach, the internal forces of each cross
section of the beam with the advance of corrosion, the actual As in each
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   43

Figure 1.21.  Corrosion levels along the length of the beam based on DSR for
(a) SC, (b) S, (c) CB, and (d) comparison of calculated corrosion levels

cross section can be calculated during the corrosion progress as well as the
corresponding corrosion levels as shown in Figure 1.22a and b, r­ espectively.
It can clearly identified from these figures that, the As values calculated for
all sections of the beam are seems to constant rather than the location of
corrosion. This deals with, all the measured strains far from the corrosion
location of the flexural steel rebar increases mainly due to the increase of
curvature. Regarding to the negative calculated values, it is resulting from
the errors in the measured compression strains with the advance of corro-
sion in this stage. Eventually it can confirm that, by applying this method
the effect curvature can be separated from effect of decrease in area steel
in a certain cross section.
In sum, the calculated corrosion degrees based on CSR, DSR, and
SFM were compared with the weight loss of external model every for
4 hours of corrosion time as shown in Figure 1.23. The weight loss of the
external model calculated represented by (% As) of the area of 2 bars with
diameter of 12.7 mm.
It was found that, there is a good agreement between the measured
corrosion levels evaluated based on DSR, SFM, and actual weight loss
44  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.22.  (a) the calculated As values of each location with corrosion time,
(b) corrosion levels

Figure 1.23.  Comparison between corrosion levels calculated from


the three approaches with the weight loss of external model

of corrosion model. The small variations between the model and the cal-
culated corrosion degrees based on strain measurement related to the
efficiency of corrosion and the difficulty of effective cleaning of the
corroded part during the corrosion process. The calculated values based
on CSR method have divergent values than the others which have been
expected as a result of curvature variations with the advance of localized
corrosion.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   45

1.5.3  DAMAGE IDENTIFICATION TECHNIQUES

The ability to promptly detect, localize and quantify structural damage has
become an increasingly important factor in maintaining performance, reli-
ability and cost effectiveness in civil, mechanical and aerospace communities.
Changes in structural dynamic parameters have been extensively utilized as
an effective tool for structural integrity assessment and damage identification.
The premise for these techniques is that damage causes a change in the physi-
cal properties of the structure, mainly in stiffness and damping at the damaged
locations. These changes in structural properties in turn alter the dynamic re-
sponse behavior of the structure from its initial healthy state. Therefore, mon-
itoring of the changes in structural response parameters can be an essential
tool for the assessment of structural integrity at the earliest possible stage.
During the past few decades, several model-free vibration-based dam-
age identification methods have been developed utilizing various d­ ynamic
characteristics such as natural frequencies, mode shapes and their deriv-
atives, modal flexibility and its curvature, modal stiffness, modal strain
energy, FRF and their curvature, and power spectral density (PSD). These
techniques have received wide attention because they are simple, fast and
inexpensive (Abdel et al. 1999; Doebling et al. 1998). Excellent reviews on
these methods have been reported by Doebling (1996), and Carden (2004).
The extension of vibration-based damage identification (VBDI) to civil
engineering structures has received increasing attention in recent years.
The distributed-strain measurement techniques are more sensi-
tive to perturbations in vibration characteristics. At the same time in-
dependence of numerical simulation is necessary in practical SHM
­applications. In other words, many algorithms might still remain viable
tools for practical damage identification but their performance can be
enhanced by distributed-strain fiber-optic sensing measurement tech-
niques as investigated below. Some Long-gauge strain-based damage
indexes as follows:

1.5.3.1  Change in Strain Mode Shape Method

The mode shape curvature (MSC) method of damage detection proposed


by Pandey, Biswas, and Samman (1991) earlier presented in section 3.2.1
can be expressed in terms of the distributed-strain modes as

1 1
MSCi =
hm
∑ r ∆δ mir =
hm
∑ r δ mir

− δ mir , Eq. 76
46  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.5.3.2  Change in Strain Mode Shape Squared Method

The MSC squared method developed by Ho and Ewins (2000), earlier


presented in Section 3.2.2, expressed in terms of the MMS is given by

1 1
∑ r ∆δ mir ∑ r (δ mir ) − (δ mir ) ,
2 2
MSCSi = 2
= d Eq. 77
hm hm

where i and r are the indices for measurement location and mode, respec-
tively, and the superscript (d) represents damaged structure.

1.5.3.3  Strain Energy Damage Index Method

The bending strain energy-based damage index for a Bernoulli–Euler


beam according to Stubbs et al. (1992) can also be obtained in terms of
the MMSV δ mi, gauge lengths lmi, and distance when hmi from the inertia
axis to sensor location is

  δ d 
2
 M  δd 2 
 δ di 
2

∑ i =1  hm lmi  ∑ i = 1  i  lmi 
M
   mi  lmi +
m

   hmi  mi   hmi  
DI i = ∑ r βir = ∑ r   − nm,
   δ mi  M δd 
2 2 2
 M  δ mi  
+ ∑ i =1   lmi  ∑ i =1  h  lmi 
mi
   lmi
   hmi  h
 mi    mi  

Eq. 78

There is an indication of damage when the index β mi > 0. For equal


gauge lengths Lm and the distance hm, the damage index DI i  becomes

DI i = ∑ r = 1 βir = ∑ r = 1 
nm nm ((
 δd
mir ) + ∑ (δ ) ) ∑ (δ )
2 M
i =1
d 2
mir
M
i =1 mir
2
 − nm ,

((
 δ mir ) + ∑ (δ ) ) ∑ (δ )
2 M
i =1 mir
2 M
i =1
d
mir
2


Eq. 79

1.5.3.4 MMS Flexibility Damage Indices

∑ j =1 { Fjjd }′′ − { Fjju }′′ ,


m
∆Fε′′ij = Eq. 80
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   47

T
 δ 1r   δ 1r 
  
 δ 2r   δ 2r 
nm 1   
Fe = Fε = ∑ r = 1 2 
T
ε −1 ε
    = Φ Ω Φ . Eq. 81
ωr    δ ( m − 1)r 
δ ( m − 1)r
  
 δ mr   δ mr 
  

Similar to the study of performance of damage indices based on dis-


placement measurement data, a practical condition is simulated by
introducing different levels of normally distributed random noise into
the distributed-strain mode shapes based on the assumption that various
denoising strategies exist with the potential to filter out as much as 5 per-
cent noise. However, 1 percent random noise is added into the natural
frequencies.
It has been verified from the study that the performance of the dam-
age identification techniques improves with high spatial resolution of
measurements. Hence, the robustness to noise will be investigated with
the 40-element model. Since all methods are very successful to detect and
accurately locate the middle damages, the robustness of the techniques to
measurement noise will be restricted to scenarios C2 and C3 correspond-
ing to single and double damage cases.
Figure 1.24 (I) shows the plots of change in modal curvature, change
in modal curvature squared, and the strain energy damage index for lo-
calization of damage using exact and 5 percent noise polluted strain
mode shapes. The damage indices using acceleration from the model
successfully located damaged elements in all the damage scenarios in
the noise-free case; however, as plotted in Figure 1.24 (II), the accelera-
tion-based indexes didn’t work well in the noisy conditions.
Unlike the poor performance under displacement mode shape meas-
urements as presented in Figures 1.24, Figure 1.24 shows that the modal
curvature, the modal curvature squared, as well as the strain energy dam-
age index methods based on the distributed-strain measurement data pol-
luted with noise are still very reliable damage identification algorithms.
It is obvious that the distributed-strain sensing technique possesses many
potential merits over the traditional sensing transducers. This further re-
veals that the adverse effect of central difference method on modal cur-
vature estimations on the performance of the curvature-based damage
identification algorithms is very significant. Long-gauge FBG sensing
technique has proven to be a viable solution to the challenges confronting
civil SHM.
48  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.24.  Modal curvature-based damage identification techniques

The strain energy-based damage index also affirms the robustness


of the MMS measurement technique. In fact, apart from the sensitivity to
damage, it is evident from Figure 1.24(c) that the strain energy damage
indices are more robust to random measurement errors than the other two
methods. The huge level of noise does not affect the ability of the strain
energy method to precisely localize damages in appropriate elements of
damage scenario C5. It can be inferred that the damage index method is
also very reliable for damage identification of civil structures if MMS
measurement data are used. The results are at variance with the techniques
based on displacement measurements as shown in Figure 1.24. The fact
that the time-series and frequency responses of the measured macro-strain
versus displacement responses share the same mapping relation (Li and Wu
2007, 2008) has further strengthened the reliability of distributed-strain
measurement data for monitoring of practical civil structures.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   49

It can be concluded from this study that long-gauge distributed-strain


measurements from FBG sensors are a much more efficient choice over
the traditional measurement techniques for reliable civil SHM. It may
therefore be concluded that the performance of some algorithms might be
improved for application in civil infrastructure by using distributed-strain
fiber-optic sensing measurement techniques.

1.5.4  DAMAGE FINGERPRINTS BASED ON LONG-GAUGE


STRAIN MODAL VECTOR

1.5.4.1  Damage Fingerprints Method

Consider a beam-like structure with two local DOFs (one for vertical
translation and the other for rotation) at each node. Suppose M long-gauge
FBG sensors are installed onto the bottom surface of the beam in a dis-
tributed manner. As described in our previous study (Li and Wu 2007), an
important feature named by MMSV can be defined as:

{δ1r , δ 2 r ,..., δ mr ,...δ Mr }T , Eq. 82

where the component corresponding to the mth sensor may be written by:

δ mr =
hm
Lm
(
φir − φ jr , ) Eq. 83

where hm is the distance from the inertia axis to the bottom of the beam; Lm
is the sensor gauge length; jlr is the mode shape at the lth DOF concerning
the r th mode.
It has been proved that the MMSV versus mode shape and the re-
sponses of the measured macro-strain versus displacement responses
share the same mapping relation. In other words, MMSV is a parameter
similar to curvature mode shape. Furthermore, MMSV only emphasizes
the relative ratio of all components. But in many cases, such as ambient
vibration under operational conditions, it is often difficult to normalize
MMSV before and after damage under the identical criteria, for example,
mass-normalization. It is necessary to provide an index free from normal-
ized criteria and suitable for any MMSV of intact and damaged structure.
An index in terms of modal strain energy is hence derived for locating
the damage (Li and Wu 2007). However, there are some limitations for
this method, such as the facts that the precision of the extracted MMSV
is not good enough and damage quantification has to still rely on the
50  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

construction of MMSV among the distributed fiber-optic sensors to de-


velop a method for both locating the damage and quantifying it.
Consider the MMSV as shown in Eq. 82. Taking a component (i.e.,
the MMS from a given sensor) from the bth sensor denoted by dbr for
­reference, the MMSV can be normalized as
δ δ δ 
{ψ 1r ,ψ 2 r ,,ψ mr ,} =  δ 1r , δ 2 r ,, δmr , . Eq. 84
 br br br 
Another damage index vector can be hence defined by the relative
error of the normalized MMSV from intact and damaged structures as

{β1r , β 2 r ,, β mr ,} , Eq. 85

with each component obtained by,


ψ mr
*
− ψ mr
β mr = × 100% , Eq. 86
ψ mr
where the superscript “*” represents the damaged state. An assumption is
made that the index from any order mode as shown in Eq. (85) can provide
an indication to damage location and quantification, which will be veri-
fied by the following numerical and experimental investigations.

1.5.4.2  Implementation Process

The aforementioned concept of damage detection will be applied to


SHM considering that the target structure is instrumented with a series of
long-gauge distributed sensors. Application of the proposed damage iden-
tification strategy to SHM is presented schematically in Figure 1.25. It is
obvious from Figure 1.25 that the proposed SHM strategy not only identi-
fies or localizes the damage in a structure but also provides additional ad-
vantages, such as: (a) the linear regression analysis provides two features:
the slope that is used to decide about the health condition of the structure,
and the R2 value that can help assess the quality of the data sensed by
different sensors; (b) the graphical user interface can help automatically
assess the conditions of different sensors at any time, that is, the reliability
of the sensing system can be evaluated easily; (c) in a continuous SHM
scheme, handling the large volume of data sensed by various sensors is a
big challenge. The proposed SHM strategy deploys some model developed
from a reasonable amount of observed data from the monitored structure.
Therefore, the data that does not reflect any behavioral change of the mon-
itored structure can easily be archived and the data load on the monitoring
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   51

Figure 1.25.  Flowchart of SHM using MMSR approach

system can be reduced; (d) the proposed SHM strategy deploys only one
type of sensor, therefore, it reduces the handling of many different types
of interrogation systems and makes the monitoring process convenient.

1.5.5  BEARING DEGRADATION DETECTION

The change of bearing status during the process of structural service is


a category of structural damage. The main function of the bearing in the
earthquake is to limit the displacement of the horizontal girders and to
provide support for the girder. The displacement of the girder is the der-
ivation of the bearing damage. The falling girder is the extreme situation
of the girder displacement. Bearing damage can be the key factor for the
loss of bearing capacity of simply supported beams. For a steel bridge,
52  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

this is a serious problem since the metal bearing in the humid environment
is susceptible to corrosion effect, or because the change of the restriction
caused by the bolt off and other reasons in the structure will change the
mechanical properties and even endanger the safety of the bridge. The
main changes of the bearing include abnormal deformation of bearing
parts in steel bridges or bolt off and the bearing parts of steel bridge suf-
fering from severe corrosion and other reasons.
How to timely identify the bearing damage in a structure is an en-
gineering problem urgently needed to be solved in the structural health
monitoring field. The damage fingerprinting method based on the statis-
tical feature of the long-gauge strain vector is also suitable for the bear-
ing damage identification of the structure. The main idea is to extract the
strain mode for multiple times and select a unit as a reference unit. The
other unit can be set as the target unit. The reference unit strain mode value
can be established for the abscissa, the target unit strain mode value for the
vertical axis. We can then draw a number of scatter dots, and then use the
straight line to draft a scatter diagram to observe the changes of the slope
of the fitting line. If the two periods of the data collected through multiple
times show the obvious change of the slope of the fitting line, the change
of the slope can determine whether there is damage to the bearing.
In the aforementioned International Bridge Study (IBS) bridge, as an
example to identify changes in the bearing, since a destructive test is not
allowed to be performed on the investigated steel string bridge, an FE
model of the southbound span 2 of the bridge is constructed in the SAP
software as shown in Figure 1.26 to demonstrate structural damage detec-
tion by the MMS ratio. The model is a combination of beam elements and
shell elements used to simulate the steel girders and the concrete deck. The
model has the same skew as the tested bridge. The concrete is modeled as a
plate with a constant thickness of 24 cm. The concrete material is defined

Figure 1.26.  The FE model of the southbound span 2


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   53

with an elastic modulus of 2.1  ×  104 MPa and a density of 2,400  kg/m 3 .
Eight girders are modeled with beam elements whose d­ imensions are the
same as those of the tested real bridges. The FE model is simply supported
at either end. It includes a total of 516 nodes, 256 frame elements, and
512 shell elements.
The ambient forces are used to excite the structure, and dynamic
analyses are performed to output dynamic structural responses. From the
calculated rotation time histories at two ends of a beam element, ­macro-
h(θ A (t ) − θ B (t ))
strain time histories are simulated by using ε (t ) = , where
L
h = the distance from the beam bottom to the neutral axis, L = e­ lement
length, θ A (t ) and θ B (t ) are rotations at two ends. The FE model simulating
the ­intact model is analyzed six times with different ambient excitation
to output six dynamic macro-strain data sets; 10 percent white noise is
added into the simulated data to represent the observation noise, in which
10 ­percent means that the standard deviation of the noise is 10 percent of
that of the simulated macro-strains.
Steel girders of bridges are generally simply supported structures with
pin and rocker bearings, which allow the girder to freely rotate. However, due
to corrosion the boundary condition may be changed to be fully or partially
fixed. This changes the structural mechanics, thus threatening the structural
safety. To simulate this kind of boundary damage, the boundary condition
of the left end of girder 3 is changed to a fixed end from a pin end in the FE
model. After calculating the dynamic responses of the “damaged” structure
and performing spectral analyses of the macro-strain time series, the MMS
ratio between the reference sensor 17 and the target sensor 1 is plotted as
shown in Figure 1.27. Similarly, the figures taking the target sensors 5 and
19 are also plotted in Figure 1.27. It can be seen that when the target sensor is
closer to the “damaged” boundary, the slope difference is larger. Figure 1.28

1 2 3
1 2 3
Tar

get

poi
Reference point

Boundary damage Indirect recognition Main beam dynamic index change mode

Difficult to directly monitor Easy to monitor

Figure 1.27.  Damage detection in the boundary damage case


54  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.28.  Slope ratio feature in the


boundary damage case

further verifies this feature, in which the “slope ratio” denotes the ratio of the
slope of the regression line of the damage structure to that of the intact struc-
ture. Therefore, this feature can be used to identify the boundary damage.
Namely, when the slope ratio progressively increases when the target sensor
is closer to a boundary, it indicates the boundary is damaged.

1.5.6  DAMAGE IDENTIFICATION OF STRUCTURES UNDER


MOVING LOADS

This section aims to develop a novel damage assessment method of bridges


under moving vehicular loads based on long-gauge strain response. The
presented method is based on the strain response when the vehicle passes
by the bridge. Numerical simulation, lab experiment, and field testing of
a real bridge have been conducted to verify the method. Moving vehicular
load is applied in this method so that the diagnosis can be carried out faster
compared to static load testing. The macro-strain-based damage index is
very sensitive to local damage covered by the gauge length of long-gauge
strain sensors, and less influenced by measurement error compared with
acceleration-based method. Meanwhile, damage index from long-gauge
stain sensors can overcome the drawback of that from traditional “point”
strain sensor (mentioned earlier).

1.5.6.1  Theoretical Background

For a beam-like bridge (Figure 1.29), the span and the height of the struc-
ture are l and H, respectively. The coordinates of sections j and j + 1 are xj
and xj+1, respectively. The distance between adjacent sections is L.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   55

Figure 1.29.  Beam-like bridge under moving vehicular load

The average moment influence line of the zone between sections j and
j + 1 can be expressed as
x j +1 x (l − Z )  xj L
M j , j +1 ( x ) = ∫x M ( x, z )dz / L = dz = x  1 − −  0 ≤ x< x j
j lL  l 2l 
x j +1 x z (l − x) x j +1 x ( l − z )
M j , j + 1 ( x ) = ∫ M ( x, z )dz / L = ∫ dz + ∫ dz
xj xj lL x lL
2 lxx j + 2 lLx − lx − 2 xLx j − L x − lx j
2 2 2
= x j ≤ x < x j +1
2 lL
x j +1
M j , j + 1 ( x ) = ∫ M ( x, z )dz / L = ∫
x j +1 z ( l − x )
dz =
(
(l − x) 2 x j + L ) x j +1 ≤ x < l,
xj xj lL 2l

Eq. 87

where M(x, z) is the moment influence line at the coordinate z, x is the


coordinate of unit vertical load P.
The physical equation is expressed as

M ε Eq. 88
= ,
EI y
where M , ε , EI , and y are the moment, strain, bending stiffness, and neu-
tral axis height in an average concept, respectively.
Substituting Eq. 87 into Eq. 88, the average strain influence line of the
zone between sections j and j + 1 can be expressed as
56  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

 xj L
x 1 − − y
 l 2l 
f j , j +1 ( x ) = 0 ≤ x < xj
( EI ) j , j + 1

f j , j +1 ( x ) =
( 2lxx j + 2 lLx − lx 2 − 2 xLx j − L2 x − lx 2j y ) x j ≤ x ≤ x j +1
2 lL( EI ) j , j + 1
( l − x )(2 x j + L ) y
f j , j +1 ( x ) = x j +1 < x ≤ l,
2 l ( EI ) j , j + 1

Eq. 89

where ( EI ) j , j + 1 is the average bending stiffness within the zone b­ etween


sections j and j + 1.
The passing vehicle (Figure 1.29b) has a total of n axles, and each
axle has the force amplitude of Pi (i = 1, 2, .  .  ., n). When the vehicle
moves across the structure at the speed v, the average strain of the zone
between sections j and j + 1 can be expressed as
n
ε j , j +1 ( x ) = ∑ Pi f j, j +1 ( x − di ) , Eq. 90
i =1

where di denotes the distance between the ith axle and the first axle, and
d1 is defined as zero. x denotes the distance between the first axle and the
left bridge support.
Eq. 90 can be further expressed by an integral along the length direc-
tion of the structure:
n n
l + dn l + dn l
∫0 ε j , j+1 ( x ) dx = ∑ Pi ∫0 f j , j + 1 ( x − di ) dx = ∑ Pi ∫0 f j, j +1 ( x ) dx,
i =1 i =1

Eq. 91
l
where ∫0 f j, j +1 ( x ) dx is the area enclosed by x axis and the average strain
influence line fj,j+1.
From Eq. 89, the area of average strain influence line can be expressed as

 l − x − 1 L  x 2 y + 2 x + L  l − lx + x j + 1  y
2 2

l  j
2 
j j (
 2 j +)1
2 
∫0 f j , j +1 ( x ) d x =
2 l ( EI ) j , j +1

( 2lx j
L
2
) l
(
+ 2 lL − 2 Lx j − L2  x j +  y − x 2j + 1 + x 2j + x j + 1 x j y
 3
)
+
2 l ( EI ) j , j + 1
g ( xi , l, L, y )
=
( EI ) j , j + 1

Eq. 92
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   57

where g(xi, l, L, y) is a function of location, length, average neutral axis


height, and boundary condition. The average strain influence line area is
only in relation to the local stiffness ( EI ) j , j + 1 of the structure and it is
­independent of the external load.
Considering the relationship between the distance x and speed v,
Eq. 91 can be transformed into

n
l + dn tn l
∫0 ε j , j +1 ( x ) dx = υ ∫ ε j , j +1 ( t ) dt =
t0
∑ Pi ∫0 f j, j +1 ( x )dx , Eq. 93
i =1

where ej, j + 1(t) is the average strain time history of the zone between sec-
tions j and j + 1. t0 is the time when the first axle enters the structure, while
tn
tn is the time when the nth axle leaves the structure. ∫t ε i (t ) = Ai (t ) is the
0

area enclosed by time axis and the average strain time history, and it can
be denoted as Aj.
Substituting Eq. 92 into Eq. 93 results in
n

tn
(
g x j , l , L, y )∑ P j

ν ∫ ε j , j +1 ( t ) dt = ν Aj =
j =1 . Eq. 94
t0
( EI ) j , j +1
Let us assume the bottom of the zone between sections j and j + 1 is
installed with jth long-gauge strain sensor, the area of strain time history
from this sensor can be obtained using Eq. 94.
n
(
g x j , l , L, y )∑ P i
i =1 . Eq. 95
Aj =
ν ( EI ) j , j +1

In a similar way, the area of strain time history from a reference ­sensor
Ar can be obtained:
n
g ( xr , l, L, y ) ∑ Pi
Ar = i =1 . Eq. 96
ν ( EI )r , r +1

The strain time history area ratio (STHAR), which is defined as the
ratio of the area of strain time history from a target sensor to that from the
reference sensor, is as follows:

β jr =
Aj
=
( )
g x j , l, L, y ( EI )r , r + 1
. Eq. 97
Ar g ( xr , l, L, y ) ( EI )r , r + 1
58  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

A total of m long-gauge strain sensors are supposed to be arranged


and the STHAR vector can be constructed when no damage occurs in the
structure:

} =  AA , AA  A  AA 
Aj
{β 1r , β 2 r  β jr  β mr
1 2 m
 , Eq. 98
 r r r r 
When damage emerges within the area covered by the jth sensor, the
STHAR from jth sensor can be written as
*
 Aj  g ( x j , l, L, y )( EI )r ,r + 1
β *jr =   = *
, Eq. 99
 Ar  g ( xr , l, L, y )( EI ) j , j + 1
*
where ( EI ) j , j + 1 is the average remaining bending stiffness within the zone
between sections j and j + 1 after damage occurrence, and it is equal to
(1 − α j )( EI ) j , j + 1 . aj is the damage extent. In this state, the STHAR vector
can therefore be changed as
 A1 A2  Aj  * Am 
{ β1r , β 2 r ...β *jr ...β mr } =  , ...   ...
A A  Ar  Ar 
, Eq. 100
 r r 

From Eq. 97, it is obvious that STHAR is independent of vehicular


weight, vehicular velocity (different measurement samples), and axle
numbers of vehicular load and is only related to the local bending stiff-
ness. When no local damage occurs, the STHAR vector will remain the
same under different tests. However, when local damage occurs within
the zone covered by a certain sensor, the STHAR from this sensor will
increase, and the STHAR vector will have a sudden change. So the
­STHAR vector can be selected as damage index. A total of mth STHAR
vectors can be calculated by successively selecting each sensor as the
reference sensor. If boundary condition changes (global damage), all
the values from these STHAR vectors except those from reference sen-
sors will vary simultaneously. This feature can be used to judge whether
global damage occurs. It should be stressed that the vehicular velocity
can be different for different measurement samples, but the vehicular
velocity should remain the same within one certain measurement sam-
ple (based on Eqs. 93–97). In practical engineering applications, only
several seconds are needed for the vehicle to pass by the bridge, and the
vehicular velocity can be approximately considered as the same during
each sample.
The damage extent aj can therefore be identified using Eqs. 97 and 99.
*
( EI ) j , j + 1 β jr
αj = 1− = 1− . Eq. 101
( EI ) j , j + 1 β *jr
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   59

Based on the above derivation process, the damage index STHAR


will fail if no strain sensor covers the damage position. In practical engi-
neering, the critical areas of beam-like structures (such as the middle area
of simply supported bridges) should be arranged with long-gauge strain
sensors. So the damage can be identified even if damage occurs randomly
within the critical areas. It is reasonable to select the reference point where
the possibility of damage occurrence is relatively small. The method of
selecting the reference sensor can be found in Hong et al. (2012).
Additionally, a reference-free method is introduced. Suppose a total
of n sensors are attached on the structure, the area of strain time history jth
sensor of initial state is A0j , and it can be expressed using Eq. 95.
n
(
g x j , l , L, y )∑ P i
0

Aoj = i =1 , Eq. 102


0
ν 0 ( EI ) j , j + 1

Similarly, the area of strain time history jth sensor of another state is A1j.
n
(
g x j , l , L, y )∑ P i
1

i =1 , Eq. 103
A1j = 1
ν ( EI ) j , j + 1
1

0 1
where ( EI ) j , j + 1 and ( EI ) j , j + 1 are the a­ verage bending stiffness within the
zone between sections j and j + 1 for the initial state and another state,
n n
respectively. ∑ Pi0 and ∑ Pi1 are the total weight of the vehicle for the
i =1 i =1
initial state and another state, respectively. v0 and v1 are the speed of
vehicle for the initial state and another state, respectively.
The ratio of A1j to A0j is defined as g10
jj  and can be expressed as

n
( EI ) j , j +1ν 0 ∑ Pi1
0

A1j i =1
γ 10
jj = n
. Eq. 104
A0j
∑ Pi
1
( EI ) j , j +1ν 1 0

i =1

The reference-free damage index can be defined as

 1
A21 A1j An1 
{γ 11 , γ 22 γ jj γ nn
10 10 10 10
} =  AA 1
0
, 
A20 A0j
  .
An0 
Eq. 105
 1

When no damage occurs within the area covered by all sensors, the
following vector can be obtained:
60  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

 0 n 1 0 n 1 n n
1
 ν ∑ Pj ν ∑ Pj ν ∑ Pj ν ∑ Pj 
0 1 0

 j =1 .
{γ 11 , γ 22 γ jj γ nn
10 10 10 10
} =  n ,
j =1
n  n
j =1
 n
j =1

 ν 1 ∑ Pj0 ν 1 ∑ Pj0 ν 1 ∑ Pj0 ν 1 ∑ Pj0 
 j =1 
 j =1 j =1 j =1 
Eq. 106

When damage occurs within the area covered by jth sensor, this vector
can be obtained:

 0 n 1 0 n 1 n n
1
 ν ∑ Pj ν ∑ Pj ( EI ) j , j + 1ν ∑ Pj ν ∑ Pj 
0
0 1 0

 j =1 .
{γ 11 , γ 22 γ jj γ nn
10 10 10 10
} =  n ,
j =1
n 
∗1
j =1
n  n
j =1

 ν 1 ∑ Pj0 ν 1 ∑ Pj0 ( EI ) ν 1
∑ Pj
0
ν 1
∑ Pj
0 
j , j +1
 j =1 
 j =1 j =1 j =1 
Eq. 107

From Eqs. 106 and 107, all the ratios g10jj (i = 1 to n) are equal when no
damage occurs, while g*10 jj will deviate from other ratios g10
jj (i = 1 to j − 1,
j + 1 to n) when damage occurs within the area covered by jth sensor. This
feature can be used for reference-free damage identification.
The presented damage assessment method can be extended to contin-
uous beam bridges, and the middle areas and hogging moment regions of
each span can be selected as the critical areas. For multigirder bridges, the
strain influence line of the section of each girder is related to the load loca-
tion along the width of the bridge based on Eqs. 87 and 89. Therefore, the
vehicular loads on such kinds of bridges should be in the same traffic lane
for different measurement samples. In addition, measurement sample is ef-
fective for damage identification only if single vehicle passes by a bridge
or several vehicles with similar speed pass by a bridge. For bridges with
small traffic flow, operational vehicular loads can be selected for damage
identification. In addition, the standard vehicle for which the weight and
configuration are known can also be arranged for damage identification.
So both rapid diagnosis and long-term SHM can be conducted for this kind
of bridge. For railway bridges, the operational train load can be selected for
damage identification. Both rapid diagnosis and long-term SHM can be
conducted for railway bridges. For bridges with large traffic flow, it is hard
to select single operational vehicular load moving on the bridge. The stan-
dard vehicle could be arranged for damage identification during the night,
when the traffic flow is relatively small. The traffic may be interrupted for
several seconds. Only rapid and periodic diagnosis can be conducted for
this kind of bridge.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   61

Figure 1.30.  The influence of vehicular velocity on the method. (a) The typical
strain time history (C0)–F6, (b) The STHAR distribution for C1–C3, (c) the
STHAR distribution for C4–C6, and (d) the STHAR distribution for C7–C9

A simply supported girder with the length of 10 m, width of 0.3 m,


and height of 0.6 m was considered for simulation. A double-axle vehicle
load with the amplitude P1 and P2 of 50 kN and speed v of 36 km/h and
72 km/h is applied for the analysis. The distance between P1 and P2 is
1 m. The typical strain time history (C0) from sensor F6 of structure under
moving vehicular load is shown in Figure 1.30a. Sensor F2 is selected
as the reference. The damage index of STHARs for all damage cases is
calculated using Eq. 97, and the results are shown in Figure 1.30b–d for
damage states-1, damage states-2, and damage states-3, respectively. It is
obvious that vehicular velocity has no influence on the method, and the
results further verify the method is independent of vehicular velocity (for
different measurement samples).

1.5.7  TWO-LEVEL DAMAGE DETECTION STRATEGY

A two-level damage detection strategy is proposed in Figure 1.31. Based


on macro-strain time-series data from distributed FBG sensors, modal
62  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.31.  Two-level damage detection strategy based on modal parameters


from dynamic macro-strain distribution

parameters including natural frequencies and MMSV can be extracted via


experimental modal analysis. Regarding MMSV, Level 1 damage location
finding approach with a space resolution of the gauge lengths of FBG
sensors can be performed with no need for a detailed analytical model.
The unknown parameters to be identified are then greatly reduced into the
damaged domain, for which frequency-based sensitivity equations may
be constructed and led to the successive Level 2 damage location finding
with a space resolution of the size of FEs and damage quantification. It
is worth stressing that “space resolution” represents the minimum region
in which it can be decided whether or not the damages have happened.
In Level 1, the space resolution depends on the gauge lengths of FBG
sensors, that is, if a damage is detected, it can be located up to the gauge
length of one FBG sensor. Regarding the FE model, in Level 2 the space
resolution depends on the sizes of FEs that can be controlled artificially
and are usually smaller than the gauge lengths of sensors.
The existing methods for the extraction of modal parameters are gen-
erally classified as either a time-domain or frequency-domain method.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   63

Regarding different dynamic testing cases, including controlled forced vibra-


tion, ambient vibration, and earthquake excitation, the corresponding strat-
egies for modal parameter extraction are diverse. Currently, the stochastic
subspace identification method seems a good choice for modal parameter
identification under ambient vibration. As the identification of modal param-
eters can be directed to the special domain on modal analysis, the following
discussion only focuses on the damage detection strategy.

1.5.7.1  MMS Vector-Based Damage Locating Method

As mentioned earlier, the dynamic macro-strain distribution, in a sense


identical to curvature measurements, directly provides a modal parameter
similar to curvature mode shape, as named by MMSV. A nonbaseline dam-
age locating method based on such MMSV has been presented in previous
research by the authors (Li and Wu 2007), for which the concept is briefly
reviewed as follows.
Suppose M long-gauge sensors are distributed throughout the beam,
an MSV of M-dimension with each component from Eq. 109 or Eq. 110
can be assembled from the displacement vector including vertical transla-
tion and rotation by left-multiplying a mapping matrix [B] as

 h1 
 L ⋅ ( vr1 − vs1 ) 
ε 1   1   v1 
   h2 ⋅ v − v  
ε 2   ( r 2 s 2 )  v 2 
MSV =   =  L2  = [ B ]M × N ⋅   , Eq. 108
      
ε M   h  v N 
 M ⋅ ( vrM − vsM ) 
 LM 
As Twp(w) (m = 1, 2, ..., M) can be measured respectively, the com-
plete set of macro-strain transfer functions may be obtained by

 v1 (ω ) 
 f (ω ) 
T1 p (ω )   p 
T (ω )   v2 (ω ) 
 2p   
{Tmp (ω )} =  =
MSV
f p (ω )
= [ B ]M × N  f p (ω )  .
     
TMp (ω )    Eq. 109
 vN (ω ) 
 f p (ω ) 
 H1 p (ω ) 
 H1 p (ω ) 
= [ B ]M × N   
 
 H Np (ω ) 
64  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

It can be verified from experimental modal analysis that for a linear


structure under single-point excitation the mode shape {jni} for the rth
mode can be represented directly from the imaginary part of measured dis-
placement transfer functions when ω = ω r . Similarly, a modal vector {dli}
for the ith mode, say MMSV, can be constructed directly from the imag-
inary part of the measured macro-strain transfer functions when ω = ω r.
Obviously from Eq. 109, it holds that

 δ 1i   ϕ1i 

 δ 
 ϕ 2 i  . Eq. 110
   = [ B ]M × N ⋅   
2i

   
δ Mi  ϕ Ni 
It should be noted that MMSV and mode shape only emphasize the
relative ratio of all components but ignore the amplitude. Furthermore, in
many cases, it is often difficult to normalize MMSV before and after dam-
age under an identical criteria, for example, mass-normalization. Hence, it
is necessary to provide an index free from normalized criteria and suitable
for any MMSV of intact and damaged structure. Here, a damage indica-
tion b is proposed by taking the original idea from Stubbs et al. (1992), who
presented a damage index method based on modal strain energy to locate
damage in structures using the measured mode shape of a few modes be-
fore and after damage.
According to Stubbs, for an N-DOF linearly elastic beam of length L
and NE elements, the following equation can be obtained:

∑ i =1 ∫a ( d 2ϕ ki* / dx 2 ) dx / ∫0 ( d 2ϕ ki* / dx 2 ) dx
N ak + Lk 2 L 2
EI k
= k
, Eq. 111
( ) ( )
a +L L
EI k*
∑ i = 1 ∫a 2 2 2 2
N
ϕ ∫0 ϕ
k k
2 2
d ki / dx dx / d ki / dx dx
k

where superscript * represents the damaged state, EIk is the flexural stiff-
ness of the kth element, and ak is the starting point of spatial coordinate of
the kth element of Lk length. It is worth noting that different from Eq. 110,
ϕ ki is the deflection mode shape for the ith mode at kth element rather than
displacement mode shape at some DOF.
On the other hand, suppose each element is installed with a long-gauge
FBG sensor, that is, M = NE. It should be stressed herein that the long-
gauge sensor is designed for macro-strain measurement so that every point
within the gauge length shares an identical strain and, hence, an identical
MMS. Therefore, the curvature mode shape can be given as

d 2ϕ ki δ Eq. 112
= ki ,
dx 2 hk
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   65

where hk is the distance from the inertia axis to the sensor location of the
kth element. Assuming hk over the beam is the same, a similar damage
evaluating index dk as in Eq. 111 can be defined:

∑ i =1 (δ ki* ) Lk / ∑ k =1(δ ki* ) Lk 


N 2 M 2

βk = − 1, k = 1, 2,..., M . Eq. 113


∑ i =1 (δ ki ) Lk / ∑ k =1(δ ki ) Lk 
N 2 M 2

The combination of all the indexes [ b1, b2, ..., bM] provides an indi-
cator to damage location. It can be seen from Eq. 111 that damage often
brings on the decrease of flexural stiffness and the fraction is larger than
the unit. So here “−1” is added to give a clearer display that bk > 0 when
there is a damage. However, as Eq. 111 is an approximate result, this con-
clusion is not always reliable and unsuitable for performing quantitatively.
The better alternative is to combine all the indexes [ b1, b2, ..., bM] to locate
damage from the positively highlighted part.

1.5.7.2 Natural Frequency-Based Damage Quantification

The frequency-based method theoretically depends on the sensitivity


equations using the relation of changes between eigenvalue and structural
stiffness and/or mass (Friswell and Mottershead 1995) as

[ S ]M × Nu ⋅ {∆p}Nu = {λ } M
T  ∂[ K 0 ] ∂[ M 0 ]  i ,
Sij = {ϕ 0i }  − λ0i  {ϕ 0 }
Eq. 114
 ∂ pj ∂ pj 

where [S] is the sensitivity matrix uniquely determined by the initial struc-
tural models. Δp is the change of object parameter. Δλ is the change of
measured eigenvalue. Before damage identification, an original FE model
should be set up and then a baseline model can be obtained after model
­updating based on Eq. 114. So the global stiffness matrix [K0] and the
mass matrix [M0] of the intact structure are obtained, from which the sen-
sitivity matrix [S] may be determined. It can be found from Eq. 114 that
once the damage denoted by Δp is located, the number of unknown pa-
rameters and, hence, the sensitivity equations are greatly reduced, which
is helpful for the convergence and reliability of the solution to the linear
equations. It is worth noting that Eq. 114 is based on first-order sensitivity
and is only suitable for small degrees of damage. If damage increases,
a second-order sensitivity analysis or step-by-step calculation should be
carried out.
66  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.6  LOAD IDENTIFICATION

The performance of bridges depends on many factors, including the type


of superstructure, support states, construction materials and quality, envi-
ronmental conditions, and external loading condition. It is important and
necessary to identify and evaluate the loads applied to bridges during their
effective service life for assessing the condition and the remaining life
of bridges. Furthermore, the influence of moving loads on the structural
response of the bridges is important to be considered for the safety of the
structure. Applying identified loads allows for a more clear understanding
as to how the bridge responds to different types of loads, although ambient
loads are helpful to assess the changes in bridge condition as monitored
over time. This section aims at developing a simple and applicable method
for the moving load identification of small and medium-sized bridges
based on the influence line theory and distributed long-gauge optical fiber
sensing technique.

1.6.1  LOAD AND SPEED IDENTIFICATION BASED ON


INFLUENCE LINE

Generally, the vehicle loads applied to a bridge can be modeled as a group


of point loads moving across the bridges. For a simply supported bridge,
the schematic of the simply supported beam is shown in Figure 1.32. Let
us take a four-axle truck moving load as a sample moving load on a simply

Figure 1.32.  Moment influence line at mid-span of a simply supported beam


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   67

supported bridge, as shown in Figure 1.32. The distances between each


axle are d1, d2, and d3 and the axle loads are P1, P2, P3, and P4, respect-
ively. The moment at the mid-span of the simple beam can be written as:

F ( x ) = P1 ⋅ f ( x ) + P2 ⋅ f ( x − d1 ) + P3 ⋅ f ( x − d1 − d2 ) + P4 ⋅ f ( x − d1 − d2 − d3 ),
Eq. 115

0 (0 ≤ x ≤ d1 )
f ( x − d1 ) =  Eq. 116
 f ( x − d1 ) ( d1 < x ≤ l + d1 ) ,

0 (0 ≤ x ≤ d1 )
f ( x − d1 − d2 ) = 
 f ( x − d1 − d2 ) ( d1 + d2 < x ≤ l + d1 + d2 ),

Eq. 117

0 (0 ≤ x ≤ d1 )
f ( x − d1 − d2 − d3 ) = 
 f ( x − d1 − d2 − d3 ) ( d1 + d2 + d3 < x ≤ l + d1 + d2 + d3 ),

Eq. 118

where x is the distance from the first axle to the left end of the bridge,
f (x) is the moment influence line at the mid-span of the bridge, and F(x) is
a function of vehicle load influence line.
This model is about a four-axle vehicle moving on the simply sup-
ported girder bridge, and it has similar expression for a multiaxle vehicle
moving on the continuous girder bridge. Gross vehicle weight (GVW) is
determined from the method of Ojio and Yamada. The strain response of
the bridge can be written as
N
ε ( x) = ∑ Pn f ( x − dn ), Eq. 119
n =1

where Pn is the weight of the nth axle, dn is the distance between the
axles, and f (x − dn) is the influence line of the continuous girder bridge.
The influence area, A, of a single truck passing over the bridge is
defined as

A( x ) = ∫−∞ ε ( x )dx Eq. 120

Substituting Eq. 119 into Eq. 120 leads to


N

A= ∑ Pn ∫−∞ f ( x − dn )dx Eq. 121
n =1
68  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

wherein
∞ ∞
∫−∞ f ( x − dn )dx = ∫−∞ f ( xn )dx = α
The integral value is a constant.
As known, the GVW is defined as
N
P= ∑ Pn Eq. 122
n =1

Consequently,

A = Pα Eq. 123

If the GVW of test truck is known, the GVW of any second truck can
be determined:
Ak Aj
α = = Eq. 124
Pk Pj

where Ak and Pk are the calculated area and reference GVW for a test
vehicle of known weight, and Aj and Pj are the calculated area and GVW
for a vehicle of unknown weight.
In fact, most small and medium-sized bridges are continuous bridges.
The studied field bridge of Xinyi River Bridge, located in Jiangsu Prov-
ince, China, is a continuous prestressed concrete bridge, which can be
simplified as a six-span with each span consisting of a 30 m continuous
beam, as shown in Figure 1.33. Consequently, such a model was used to
study the speed and load identification of continuous concrete bridges.
The cross sections from the left end of support cross section, 1/4 cross
section, 1/2 cross section, 3/4 cross section and right end of support cross
section of the second span are termed as A, B, C, D, and E cross section,
respectively. According to the influence line theory, the bending moment
influence line was obtained as shown in Figure 1.34. Assuming that the
cross section stiffness was a constant, the shape of strain influence line
was similar to that of the bending moment influence line.

v
A B C D E

x = vt
L = 30m L L L L L

Figure 1.33.  Model of a six span continuous bridge


STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   69

Figure 1.34.  Moment influence line of each cross section

The figure shows that the response mainly concentrated in the first,
second, and third span, and when the bending moment value is zero, the
vehicle is on support. Consequently, when calculating the influence area
and the speed of the vehicle, only three spans are needed.
The speed of the vehicle can be calculated as
d
v= Eq. 125
∆t
where d is the length of three spans, and Δt is the time when the vehicle
passed over the three-span bridge.

1.6.2  LOAD IDENTIFICATION FOR AN IN-FIELD CASE

Herein, we take the example of Xinyi River Bridge, which is located in


Shuyang, Jiangsu Province, China, and has two traffic lanes. The pan-
orama of the bridge is shown in Figure 1.35. The superstructure is com-
posed of partially prestressed concrete box girders. The total length of
the bridge is 2168.20 m with 72 spans, and each span has a length of
30 m. The monitoring system was installed on one span of the Xinyi River
Bridge. There are four box girders for each span. The arrangement of
long-gauge FBG sensors is shown in Figure 1.36, all of which are divided
into four channels, numbered as channel 1 to channel 4. The sensors with
a gauge length of 1.0 m were attached to the bottom of the girders with
epoxy resins, as shown in Figure 1.36.
70  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.35.  Panorama of Xinyi River Bridge

Figure 1.36.  Installation of long-gauge FBG sensors on the bottom of the girders
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   71

1.6.2.1  Establishment of Simulation Model of Xinyi River Bridge


and the Monitoring System

FE program ANSYS was used to establish the simulation model of Xinyi River
Bridge. The reinforcing rebars were simulated with Link 8 elements, concrete
was simulated with solid 45 elements, and the supports were simulated with
combined 14 elements. The connection between box beams was rigid.
The beam was regarded as a continuous structure, and the support
constraint was exerted to the beam to simulate the actual case. The main
loads included the self-weight, moving vehicle load, and prestress. ­After
the establishment of separate models for steel and concrete, coupled
equations were applied to couple the joints in order to ensure the proper
connections between the two parts. Before the installation of sensors, the
structure already supported its weight. Consequently, the measurement of
sensors was the strain increase due to the vehicle loads. In order to simu-
late this condition, only vehicle loads were added.
According to the material performance and on-site measurements, the
modulus of the concrete was E = 3.3 × 104 MPa, density 2,550 kg/m3,
Poisson’s ratio 0.167; the modulus of pavement was E = 1.2 × 103 MPa,
density 2,200 kg/m3, Poisson’s rate 0.35. The applied prestressed steel
strands were characterized as having high strength and low creep. The
standard strength Ryb, modulus, Poisson’s ratio, and coefficient of linear
expansion of the steel strand were 1,860 MPa, 0.95 × 105 MPa, 0.27,
and 1.2 × 10−5 m/°C, respectively. The longitudinal and traverse stiffness
coefficients of the support were 4 × 106 and 0.6 × 106 N/m, respectively.
The ANSYS model was established as shown in Figure 1.37, wherein alto-
gether there were 35,323 nodes and 23,224 elements. Two moving loads of
10 kN with 2.4 m apart from each other were exerted on the carriageway,
and the moving speed was 120 km/h.

Figure 1.37.  ANSYS model of the bridge


72  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.6.2.2  Load Identification Based on the Field Monitoring Data

1.6.2.2.1  Integral under Different Moving Loads


With the moving speed as the parameter, a serial of integrals (Eq. 120
and 121) for different speeds could be obtained, as shown in Table 1.2. It
is shown that the integrals were not directly related to the moving speed
when the lanes were smooth enough.

1.6.2.2.2  Processing and Calibration


We chose ten sample vehicles to calibrate the constant of α. The true weight
of the ten vehicles was acquired from the weigh-in-motion (WIM) system.
The traditional WIM system was installed about 200 m in front of the Xinyi
River Bridge. An example strain versus time plot was used to calculate the
influence area for one of the ten vehicles in the mid-span of 2# box girder,
as shown in Figure 1.38, where the shaded area is computed to get the in-
fluence area A2. The total area Asum was obtained through summing all the
1# to 4# box girder’s influence areas. According to the WIM value and the
total influence area, the constant α was obtained, as shown in Table 1.3.

Table 1.2  Integrals for different speeds

Integrals 40 km/h 60 km/h 80 km/h 100 km/h 120 km/h


1/2 span of 2# 9.81 9.75 9.82 9.70 9.77
1/2 span of 3# 8.07 8.01 8.08 7.93 8.03
1/2 span of 2# 9.15 9.11 9.16 9.03 8.37
left 2 m
1/2 span of 3# 7.51 7.47 7.52 7.37 7.47
left 2 m
1/2 span of 2# 8.74 8.66 8.74 8.62 9.10
right 2 m
1/2 span of 3# 7.11 7.04 7.11 6.97 7.20
right 2 m
Left support −12.75 −12.82 −12.75 −12.83 −13
of 2#
Right support −10.83 −10.83 −10.82 −10.89 −11
of 3#
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   73

Figure 1.38.  Strain–time curve for a typical vehicle at the mid-span of 2# box girder

Table 1.3  Calibration of constant α

Vehicle Asum WIM weight Constant α


Vehicle 1 1059.881 61.6 17.21
Vehicle 2 929.4131 53.2 17.47
Vehicle 3 1097.936 58.8 18.67
Vehicle 4 972.0112 59.6 16.31
Vehicle 5 1014.124 57.1 17.76
Vehicle 6 999.8889 60.0 16.66
Vehicle 7 735.0772 52.1 14.11
Vehicle 8 878.1233 54.7 16.05
Vehicle 9 294.7808 19.7 14.96
Vehicle 10 955.1497 60.2 15.87
Average 16.51

1.6.2.2.3  Moving Speed Identification


With the above method, the moving speed was identified. Table 1.4 shows
the speed identification of one vehicle with different sensors. The identified
speeds agreed well with the WIM ones, the relative error was smaller than
5 percent. Overall 130 samples were taken. The moving speed was identi-
fied with the long-gauge FBG sensor located at the 1/2 span of 2# girder
74  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

(Figure 1.39). Among the 130 samples, the relative error of 116 samples
was within 5 percent.

Table 1.4  Speed identification for one sample

Identified WIM
Location speed speed Absolute Relative
Sensor of sensors (km/h) (km/h) error error (%)
1 1/2 span 63.90 67 3.10 4.62
of 3#
2 1/2 span 64.48 67 2.52 3.76
of 1#
3 3/4 span 67.35 67 −0.35 −0.53
of 4#
4 3/4 span 67.84 67 −0.84 −1.26
of 2#
5 1/2 span 64.34 67 2.66 3.97
of 2#
6 1/2 span 65.81 67 1.19 1.77
of 4#

Figure 1.39.  Relative error for the seed identification (2# girder)
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   75

1.6.2.2.4 Load Identification


Based on the speed identification, the influence lines could be obtained
from the strain–time curves. Then, the moving loads of 50 samples were
identified with the long-gauge FBG sensor located at 1/2 span of 2# girder
according to the values in Tables 1.2 and 1.3 and the proposed methods.
All corresponding 10 results are shown in Table 1.5.
Comparing the identified moving loads with the WIM weights, the
relative error was obtained. The relative error is shown in Figure 1.40. The
relative error of moving weights was within 15 percent and most errors
were within 10 percent. The method for the moving load identification
of bridges based on the influence line theory and distributed optical fiber
sensing technique was, thus, confirmed to be effective. The error of load
identification was mainly due to the following: (1) there exists some error
in the WIM weight system, although the loads were obtained from WIM;
(2) the constants of α of each box girder may be different since the bridge
has served for more than 10 years and the bending stiffness and other
properties for each box girder are somewhat different; (3) the influence

Table 1.5  Load identification results of 10 samples (sensor located at


1/2 span of 2# girder)

Measure Simu-
integrals lated
under integrals
actual under Identified WIM Relative
Sample loads unit load load (ton) load (ton) error (%)
1 532.53 9.77 54.5 52.4 −4.0
2 611.58 9.77 62.6 61.6 −1.6
3 519.50 9.77 53.2 53.2 0.1
4 564.07 9.77 57.7 58.8 1.8
5 562.80 9.77 57.6 59.6 3.3
6 556.97 9.77 57.0 57.1 0.2
7 557.64 9.77 57.1 60.0 4.9
8 423.73 9.77 43.4 52.1 16.8
9 550.52 9.77 56.3 54.7 −3.0
10 166.37 9.77 17.0 19.7 13.6
76  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Figure 1.40.  Relative error of weight

of vibration and coupling effect of the bridge with the moving mass was
not taken into consideration. This coupling has some influence on the load
identification.

1.7  FE MODEL UPDATING METHOD

This section presents a novel strategy for FE model updating of flexural


structures. The method is based on modal parameters extracted from
distributed dynamic macro-strain responses. The objective function is
comprised of low-order MMS and frequency is established first, while
local bending stiffness, density, and boundary conditions of structures
can be selected as the design variables. The long-gauge macro-strain sen-
sors chosen in this section are firstly addressed. The fundamental sens-
ing properties of the macro-strain sensors confirm an effective dynamic
measurement capacity. Notably, the updated FE model can predict local
response (MMS) and global response (frequency and displacement mode)
because the local information-sensitive index and global information-sen-
sitive index are included in the objective function. Therefore, the proposed
FE model updating strategy constitutes a new alternative in performance
assessment of flexural structures.
The potential advantages of FE model updating using distributed
macro-strain sensing techniques can be summarized as follows: (1) MMS
is very sensitive to local damage (Hong et  al. 2012; Li and Wu 2007);
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   77

(2)  MMS-based damage index is less sensitive to noise pollution


(Adewuyi et al. 2011); (3) only low-order mode is necessary to localize
damage and can, therefore, be easily extracted under ambient vibration
(Hong et  al. 2012); and (4) both local sensitive information and global
information can be captured by distributed deployment.

1.7.1  STRATEGY OF FE MODEL UPDATING

The macro-strain FRF from mth sensor excited by the dynamic load at the
pth DOF is obtained by the displacement FRF.

ε ε m (ω ) h
H mp (ω ) = = m  H ipd (ω ) − H djp (ω )  . Eq. 126
f p (ω ) Lm
The absolute value at the peak for the rth mode can be written as
ϕ pr

ε
r H mp (ω = ωr ) = δ mr Eq. 127
2 M r ξ r ω r2

where jpr, Mr, xr, and wr represent rth displacement mode shape at the
pth DOF, rth modal mass, rth damping ratio, and rth natural frequency,
­respectively. For a given mode, jpr  / 2Mr xr w2r remains a constant. The com-
bination of macro-strain magnitude FRFs extracted from the distributed
macro-strain responses can construct a vector as dr = {d1r, d2r, . . . dmr,. . .}T.
The MMS from the target sensor can be normalized by that of the refer-
ence sensor and can, therefore, be expressed as follows:

δ δ
δ r =  1r , 2 r ,...1,...}
T
Eq. 128
δ δ
 mr mr

The physical equation can approximately be expressed as

M ir δ
= ir Eq. 129
( EI )i yir

where M ir , yir , and (EI )i represent average modal moment, average


modal neutral axis height, and average modal flexural rigidity within the
gauge length of the ith sensor, respectively. The subscript r is assigned
to the rth mode. dir is defined as MMS corresponding to ith long-gauge
­macro-strain sensor. It is obvious from Eq. 129 that the ith local bending
stiffness is directly relevant to MMS from ith long-gauge macro-strain
sensor.
78  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

An objective function that consists of both local information and


global information can be constructed considering the sensitivity of MMS
to local damage, and the capacity of frequency can reflect the global
­behavior of the structure.
n
δ ai − δ ti m
f ( x) = ∑γ i δ ti
+ ∑ β j (1 − f aj / f tj )2 Eq. 130
i =1 j =1

where dai and faj represent the ith MMSV and the jth natural frequency
from the FE method. dti and ftj represent the ith MMSV and the jth natu-
ral frequency from field measurements. gi and bj are weight coefficients.
The principle of determination of weight coefficients in this section is
g1 = 1, and the values of b1 and b2 basically satisfy the MMS residual and
­frequency residual with the same order of magnitude. || || is the Euclid
norm of a vector. In practical engineering, only low-order modal parame-
ters are necessary, which can be easily extracted under ambient vibration
or seismic excitation. It is worth noting that the reference sensor can be
selected randomly and this characteristic of the process will be discussed
in the following sections.
FE model updating is transferred to minimize the objective function
under constrained conditions and the mathematical model can be written as

 n
δ −δ m

 min f ( x ) = min ∑ γ i ai δ ti + ∑ β j (1 − f aj / f tj )2
 i =1 ti j =1
 s.t . g1 ≤ gi ≤ g2 Eq. 131
 h1 ≤ hi ≤ h2

 k1 ≤ ki ≤ k2

where g, h, and k are design variables, with g representing average bend-


ing stiffness within the gauge length of a certain long-gauge sensor, while
h and k represent mass density and the boundary condition of the struc-
ture, respectively. The first-order optimization method is applied to solve
the mathematical model in this chapter. The flowchart of the FE model
updating strategy is illustrated in Figure 1.41.
To study the effectiveness of the FE model updating method, an
I-shaped beam with 5.6 m span, 100 mm width, and 100 mm height was
designed for simulation. The identified strain mode and displacement
mode are illustrated in Figures 1.42 and 1.43, respectively. The MMS from
target sensor is also normalized with the reference sensor F6. The results
show that the updated mode shapes are roughly the same as those of the
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   79

Figure 1.41.  Flowchart of the FE model updating strategy

actual ones. If no damage occurs within the reference sensor, the damage
can be detected by change of MMS (MMS-based damage index). It can be
seen from Figure 1.42 that the MMS-based damage index fails to identify
damage occurring in E6 and E8 because damage occurs within the refer-
ence sensor F6.
80  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

1.5 1.5

1.2 Before updating


Actual-C4
1.2 0.9
After updating-C4
0.6 Actual-C5
i1

i2
After updating-C5
0.9 0.3 Actual-C6
Modal macro-strain

Modal macro-strain
After updating-C6
0.0
Before updating
0.6 -0.3
Actual-C4
After updating-C4 -0.6
Actual-C5
0.3 After updating-C5 -0.9
Actual-C6 -1.2
After updating-C6
0.0 -1.5
E1 E2 E3 E4 E5 E6 E7 E8 E9 E10 E11E12E13 E14E15 E16E17 E1 E2 E3 E4 E5 E6 E7 E8 E9 E10 E11 E12 E13E14 E15 E16 E17

Element Element

Figure 1.42.  Normalized first-order MMS and second-order MMS

1.2 1.5
1.2 Before updating
1.0 Actual-C4
x1

x2

0.9 After updating-C4


Displacement mode shape
Displacement mode shape

0.6 Actual-C5
0.8 After updating-C5
0.3 Actual-C6
Before updating After updating-C6
0.6 0.0
Actual-C4
After updating-C4 -0.3
0.4 Actual-C5 -0.6
After updating-C5
Actual-C6 -0.9
0.2
After updating-C6
-1.2
0.0 -1.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
Distance (m) Distance (m)

Figure 1.43.  First-order displacement mode and second-order displacement mode

REFERENCES

Abdel, W. M., G. De, G. Roeck, and B. Peeters. 1999. “On the Application of FE
Model Updating to Damaged Concrete Beams.” In 2nd International Confer-
ence on Identification in Engineering Systems, March, 1999, 578–587. Wales,
UK: University of Wales.
Adewuyi, A. P., and Z. S. Wu. 2011. “Modal Macro-Strain Flexibility Methods for
Damage Localization in Flexural Structures Using Long-Gage FBG Sensors.”
Structural Control and Health Monitoring 18, no. 3, pp. 341–60.
Carden, E. P. 2004. “Vibration Based Condition Monitoring: A Review.” Structural
Health Monitoring 3, no. 4, pp. 355–77.
Doebling, S. W., C. R. Farrar, M. B. Prime, and D. W. Shevitz. 1998. “A Review of
Damage Identification Methods that Examine Changes in Dynamics Proper-
ties.” Shock and Vibration Digest 30, no. 2, pp. 91–180.
Doebling, S. W. 1996. “Minimum-Rank Optimal Update of Elemental Stiffness Param-
eters for Structural Damage Identification.” AIAA Journal 34, no. 12, pp. 2615–21.
STRUCTURAL IDENTIFICATION AND DAMAGE DETECTION  •   81

Ewins, D. J. 2000. “Basics and State-of-the-art of Modal Testing.” Sadhana 25,


no. 3, pp. 207–20.
Fujino, Y., M. Murata, S. Okano, and M. Takeguchi. 2000. “Monitoring System
of the Akashi Kaikyo Bridge and Displacement Measurement using GPS.
SPIE’s, International Symposium on Nondestructive Evaluation and Health
Monitoring of Aging Infrastructure.” International Society for Optics and
Photonics 3995, pp. 229–36.
Geng, J., J. Wu, and X. Zhao. 2009. “Simulation of Fiber Bragg Grating Sensor
for Rebar Corrosion.” In Second International Conference on Smart Materi-
als and Nanotechnology in Engineering International Society for Optics and
Photonics, 8-11 July, 2009, edited by J. Leng, A. K. Asundi, and W. Ecke,
74931I-74931I. Weihai, China: SPIE.
Hampshire, T. A., and H. Adeli. 2000. “Monitoring the Behavior of Steel Struc-
tures using Distributed Optical Fiber Sensors.” Journal of Constructional
Steel Research 53, no. 3, pp. 267–81.
Ho, Y. K., and Ewins, D. J. 2000. “On the Structural Damage Identification with
Mode Shapes.” In European COST F3 Conference on System Identification and
Structural Health Monitoring, June, 2000, edited by J. A. Güemes, pp. 677–686.
Madrid, Spain: Universidad Politécnica de Madrid.
Hong, W., Z. S. Wu, C. Q. Yang, C. F. Wan, and G. Wu. 2012. “Investigation on the
Damage Identification of Bridges using Distributed Long-Gauge Dynamic
Macro-Strain Response under Ambient Excitation.” Journal of Intelligent Ma-
terial Systems and Structures 23, no. 1, pp. 85–103.
Hong, W., C. Q. Yang, Z. S. Wu, and Y. F. Zhang. 2010. “Identification of Modal
Macro-Strain Vector Based on Distributed long-Gage FBG Sensors under
Ambient Vibration, Sensors and Smart Structures Technologies for Civil.”
Mechanical, and Aerospace Systems 7647, pp. 36–44.
Lee, J. R., C. Y. Yun, and D. J. Yoon. 2009. “A Structural Corrosion-Monitoring
Sensor based on a Pair of Prestrained Fiber Bragg Gratings.” Measurement
Science and Technology 21, no. 1, p. 017002.
Li, S. Z., and Z. S. Wu. 2008. “Modal Analysis on Macro-Strain Measurements from
Long-Gauge Fiber Optic Sensors.” Journal of Intelligent Material Systems and
Structures 19, no. 8, pp. 937–46.
Li, S. Z., and Z. S. Wu. 2007. “A Non-baseline Algorithm for Damage Locating
in Flexural Structures using Dynamic Distributed Macro-Strain Responses.”
Earthquake Engineering & Structural Dynamics 36, no. 9, pp. 1109–25.
Pandey, A. K., M. Biswas, and M. M. Samman. 1991. “Damage Detection from
Changes in Curvature Mode Shapes.” Journal of Sound Vibration 145, no. 2,
pp. 321–32.
Shen, S., Z. Wu, C. Yang, C. Wan, Y. Tang, and G. Wu. 2010. “An Improved Conju-
gated Beam Method for Deformation Monitoring with a Distributed Sensitive
Fiber Optic Sensor.” Structural Health Monitoring 9, no. 4, pp. 361–78.
Smyth, A., and M. Wu. 2007. “Multi-rate Kalman Filtering for the Data Fusion of
Displacement and Acceleration Response Measurements in Dynamic System
Monitoring.” Mechanical Systems & Signal Processing 21, no. 2, pp. 706–23.
82  •   FIBER-OPTIC SENSORS FOR HEALTH MONITORING, VOL II

Stubbs, N., J. T. Kim, and K. G. Topole. 1992. “An Efficient and Robust Algorithm
for Damage Localization in Offshore Platforms.” In ASCE 10th Structures
Congress, pp. 543–546.
Tondini, N., O. S. Bursi, A. Bonelli, and M. Fassin. 2015. “Capabilities of a ­Fiber
Bragg Grating Sensor System to Monitor the Inelastic Response of Con-
crete Sections in New Tunnel Linings Subjected to Earthquake Loading.”
Computer-Aided Civil & Infrastructure Engineering 30 no. 8, pp. 636–53.
Weidner, J., J. Prader, F. Moon, and E. Aktan. 2011. Static, Crawl, Ambient and
Forced Vibration Testing (Technical report). Philadelphia: Drexel University.
Zaki, A., H. K. Chai, D. G. Aggelis, and N. Alver. 2015. “Non-Destructive Eval-
uation for Corrosion Monitoring in Concrete: A Review and Capability of
Acoustic Emission Technique.” Sensors 15, no. 8, pp. 19069–101.
Zhang, H., and Z. Wu. 2012. “Performance Evaluation of PPP-BOTDA-based Dis-
tributed Optical Fiber Sensors.” International Journal of Distributed Sensor
Networks. https://doi.org/10.1155/2012/414692
Zhao, X., P. Gong, G. Qiao, J. Lu, X. Lv, and J. Ou. 2011. “Brillouin Corrosion Ex-
pansion Sensors for Steel Reinforced Concrete Structures using a Fiber Optic
Coil Winding Method.” Sensors 11, no. 11, pp. 10798–819.
Zheng, Z., Y. Lei, and X. Sun. 2001. “Measuring Corrosion of Steels in Concrete
via Fiber Bragg Grating Sensors—Lab Experimental Test and In-Field Appli-
cation.” In 12th Biennial International Conference on Engineering, Construc-
tion, and Operations in Challenging Environments; and Fourth NASA/ARO/
ASCE Workshop on Granular Materials in Lunar and Martian Exploration,
14-17 March, 2010, edited by G. Song, pp. 2422–2430. Honolulu, Hawaii:
­American Society of Civil Engineers.
Zheng, Z., X. Sun, and Y. Lei. 2009. “Monitoring Corrosion of Reinforcement in
Concrete Structures via Fiber Bragg Grating Sensors.” Frontiers of Mechani-
cal Engineering in China 4, no. 3, pp. 316–19.
Index

A extension under continuous beam


AASHTO. See American conditions, 11–12
Association of State Highway extension with combined action
and Transportation Officials from loads and support settle-
Accidental Limit State, 109 ments, 8–10
American Association of State under temperature variation,
Highway and Transportation 10–11
Officials (AASHTO), 107 Conjugated beam theory, 6–8
American Society of Civil Continuous beam conditions,
Engineers (ASCE), 113 extension under, 11–12
American Specification Continuous-strain ratio
AASHTO, 114 (CSR), 35
ANSYS model, 74 based corrosion identification
ASCE. See American Society of technique, 36
Civil Engineers Corrosion identification technique,
34–35
B continuous-strain ratio-based, 36
Bayesian theory, 110 distributed-strain ratio-based,
Bearing degradation detection, 36–38
51–54 of RC structures, 39–44
Bernoulli–Euler beam, 46 SFM-based corrosion identifica-
Bilinear cumulative fatigue tion technique, 38–39
damage deduction, 114–115 Cumulative fatigue damage,
Brillouin scattering sensing, 124 deterioration analysis based
British Specification BS5400, 114 on, 113–121

C D
Civil engineering structures, health Damage fingerprints method,
and safety of, 32 49–50
Conjugated beam method, 6–14 Damage identification techniques,
conjugated beam theory, 6–8 45
experimental verification of, MMS flexibility damage indices,
12–14 46–49
130  •   INDEX

Damage identification Fiber Bragg Grating (FBG)


­techniques (Continued ) sensors, 118
strain energy damage index Field monitoring data
method, 46 integral under different moving
strain mode shape method, loads, 72
change in, 45 load identification based on, 72,
strain mode shape squared 75–76
method, change in, 46 moving speed identification, 73
Deflection, 4 processing and calibration,
Deterioration process 72–73
based on cumulative fatigue Field testing, evaluating carrying
damage, 113–121 capacity based on, 107–108
mechanisms and prediction of, Finite element method (FEM), 13
111–113 First Order Reliability
DIM. See Double integration Method, 110
method First Order Second Moment
Distributed-strain ratio (DSR), 35 Method (FOSM), 110
Distributed-strain ratio-based Fourier Transform (FT), 24
corrosion identification Frequency-domain method, 62
technique, 36–38 Frequency response function
Double integration method (FRF), 16
(DIM), 4 FT. See Fourier Transform
DSR. See Distributed-strain ratio
G
E Genetic algorithms, 110
European Specification Global Positioning System
EN1993, 114 (GPS), 1
Event warning, 98 Gross vehicle weight (GVW), 67

F H
Fast Fourier Transform (FFT), 26 Hi-Tec weldable quarter-bridge
Fatigue cumulative damage strain gauge, 29
deduction, 115 I
Fatigue Limit State, 109 IBS. See International Bridge
FE model updating method, 76–77 Study (IBS) bridge
strategy of, 77–80 Influence line, load and speed
Federal Highway identification based on,
Administration, 90 66–69
Fédération internationale du béton Integration method, 5–6
(FIB), 90 International Association of
FEM. See Finite element method Structural Health
FFT. See Fast Fourier Transform Monitoring, 91
FIB. See Fédération internationale International Bridge Study (IBS)
du béton bridge, 52
INDEX  •   131

J N
Japanese Society of Civil Natural frequency-based damage
Engineering (JSCE), 87 quantification, 65
Neutral axis height identification,
L 23–24
LG-FBG sensors, 29–30 experimental example, 27–31
Linear fatigue damage, deduction new method to
of, 114 determine, 24–25
Load identification procedure to, 25–27
an in-field case, 69–76
and speed identification based on P
influence line, 66–69 Packaged fiber Bragg grating
Long-gauge strain modal theory, (PFBG) sensors, 35
18–23 Performance evaluation, methods
damage fingerprints method, of, 104–106
49–50 Performance indexes, 100
implementation process, 50–51 Performance warning, 98
macro-strain FRF, 18–20 Point-type strain modal analysis,
modal parameters extraction, 14–18
20–23 PPP-BOTDA technology, 13
Prediction methods, based on
M analogy, 117
Macro-strain-based damage experience, 117
index, 54 mathematical theoretical models,
Macro-strain FRF, 18–20 117–118
Markov Chain Monte Carlo mechanical models, 118
method, 110 probability analysis, 118
Markov process, 113 short-time tests, 117
MMS flexibility damage indices, theory of structural
46–49 reliability, 118
MMS vector-based damage Progressive Limit State, 109
locating method, 63–65
Modal Macro Strain (MMS), R
101–102 Reference-free method, 59
Modal macro-strain vector Reinforced concrete (RC)
(MMSV), 21 structures
Modal parameters extraction, corrosion of steel in, 34
20–23 health monitoring for, 33
Mode shape curvature (MSC) Residual life prediction,
method, 45 theory of, 115
Monte Carlo sampling, 110 Rotation and deflection
Moving loads identification, 4–5
damage identification of struc- conjugated beam method, 6–14
tures under, 54 integration method, 5–6
theoretical background, 54–61
132  •   INDEX

S warning indexes and evaluation


S-N curve, 114 criteria, 99–104
Section fiber model (SFM), 35 Structural health monitoring
based corrosion identification (SHM)
technique, 38–39 concluding remarks and future
Sensor monitoring device, 124 work, 123–128
Serviceability, 84 current state and future trends for
Limit State, 109 development, 85–87
SFM. See Section fiber model performance evaluation guide
SHM. See Structural health for bridges, 86–87
monitoring safety evaluation methods of
SM-130, 29 bridges, 87
Steel girders of bridges, 53 physical condition of highways
STHAR. See Strain time history and bridges directly based on,
area ratio 106–107
Strain energy damage index standardization of procedures
method, 46 for, 87–92
Strain modal analysis theory, 14 theoretical background of area
long-gauge, 18–23 distributed sensing in, 92–94
point-type, 14–18 Structural identification and
Strain mode shape method, damage detection
change in, 45 based on macro-strain measure-
Strain mode shape squared ments, 1
method, change in, 46 FE model updating method,
Strain time history area ratio 76–77
(STHAR), 57 strategy of, 77–80
Structural damage identification, load identification
32–33, 123–124 an in-field case, 69–76
bearing degradation detection, and speed identification based
51–54 on influence line, 66–69
corrosion monitoring, 34–44 neutral axis height identification,
crack width monitoring, 33–34 23–24
damage identification tech- experimental example, 27–31
niques, 45–49 new method to determine,
long-gauge strain modal vector, 24–25
damage fingerprints based procedure to, 25–27
on, 49–51 rotation and deflection identifi-
structures under moving loads, cation, 4–5
damage identification of, conjugated beam method, 6–14
54–61 integration method, 5–6
two-level damage detection strain modal analysis theory, 14
­strategy, 61–65 long-gauge strain modal theory,
Structural fault (damage detection) 18–23
analysis, 101 point-type strain modal
basic concepts, 96–99 ­analysis, 14–18
INDEX  •   133

structural damage identification, T


32–33 Temperature variation,
bearing degradation detection, applicability of CBM under,
51–54 10–11
corrosion monitoring, 34–44 Time-domain, 62
crack width monitoring, 33–34 Two-level damage detection
damage identification tech- strategy, 61–63
niques, 45–49 MMS vector-based damage lo-
long-gauge strain modal vector, cating method, 63–65
damage fingerprints based natural frequency-based damage
on, 49–51 quantification, 65
moving loads, damage identi-
fication of structures under, U
54–61 Ultimate Limit State, 109
two-level damage detection Unstable mechanism, 8
strategy, 61–65
using area-wise distributed sens- V
ing data, 2–4 VBDI. See Vibration-based
Structural Identification damage identification
Committee of the American Vibration-based damage
Society of Civil Engineers, 90 identification (VBDI), 45
Structural performance evaluation
area distributed W
­monitoring-based, 83 Warning indexes and evaluation
three-level performance-based criteria, 99–104
evaluation incorporating, Weigh-in-motion (WIM) system,
95–121 72
concept of, 83
comprehensive SHM-based X
performance evaluation, Xinyi River Bridge, 68
85–87 on field monitoring data, 72–76
overview of methods for per- load identification for in-field
formance evaluation, 83–85 case, 69–70
standardization of procedures simulation model of, 71
for SHM, 87–92
theoretical background of area Z
distributed sensing in SHM, Zhishen, Wu, 91
92–94
Structural reliability analysis,
evaluation methods based on,
109–111
OTHER TITLES IN OUR SUSTAINABLE STRUCTURAL
SYSTEMS COLLECTION
Mohammad Noori, Editor
• Numerical Structural Analysis by Steven O’Hara and Carisa H. Ramming
• A Systems Approach to Modeling Community Development Projects by Bernard Amadei
• Seismic Analysis and Design Using the Endurance Time Method, Volume I : Concepts
and Development by H.E. Estekanchi and H.A. Vafai
• Seismic Analysis and Design Using the Endurance Time Method, Volume I : Advanced
Topics and Application by H.E. Estekanchi and H.A. Vafai
• Multi-Scale Reliability and Serviceability Assessment of In-Service Long-Span Bridges
by Mohammed Noori
• Using ANSYS for Finite Element Analysis, Volume I: A Tutorial for Engineers
by Wael A. Altabey
• Using ANSYS for Finite Element Analysis, Volume II: Dynamic, Probabilistic Design
and Heat Transfer Analysis by Wael A. Altabey
• Science and Technology Diplomacy, Volume I: A Focus on the Americas with Lessons for
the World by Hassan A. Vafai and Kevin E. Lansey
• Science and Technology Diplomacy, Volume II: A Focus on the Americas with Lessons
for the World by Hassan A. Vafai and Kevin E. Lansey
• Science and Technology Diplomacy, Volume III: A Focus on the Americas with Lessons for
the World by Hassan A. Vafai and Kevin E. Lansey

Momentum Press offers over 30 collections including Aerospace, Biomedical, Civil,


Environmental, Nanomaterials, Geotechnical, and many others. We are a leading book
publisher in the field of engineering, mathematics, health, and applied sciences.

Announcing Digital Content Crafted by Librarians


Concise e-books business students need for classroom and research

Momentum Press offers digital content as authoritative treatments of advanced engineering


topics by leaders in their field. Hosted on ebrary, MP provides practitioners, researchers,
faculty, and students in engineering, science, and industry with innovative electronic content
in sensors and controls engineering, advanced energy engineering, manufacturing, and
materials science.

Momentum Press offers library-friendly terms:


• perpetual access for a one-time fee
• no subscriptions or access fees required
• unlimited concurrent usage permitted
• downloadable PDFs provided
• free MARC records included
• free trials

The Momentum Press digital library is very affordable, with no obligation to buy in future years.

For more information, please visit www.momentumpress.net/library or to set up a trial in the


US, please contact mpsales@globalepress.com.

Вам также может понравиться