You are on page 1of 11

Engineering Structures 32 (2010) 17041714

Contents lists available at ScienceDirect

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

Acoustic emission monitoring of bridges: Review and case studies


Archana Nair, C.S. Cai
Department of Civil and Environ. Engineering, Louisiana State University, Baton Rouge, LA 70803, United States

article

info

Article history:
Received 12 January 2008
Received in revised form
11 December 2009
Accepted 3 February 2010
Available online 6 March 2010
Keywords:
Acoustic emission
Bridge monitoring
Bridge test
Non destructive evaluation

abstract
This paper gives a brief review of the acoustic emission technique and its applications to bridge health
monitoring. Emphasis is given to the discussion of available techniques of AE data processing, both
qualitative and quantitative. An assessment of the statistical quantitative analysis technique, intensity
analysis, is illustrated through two case studies. This technique of damage quantification is applied to AE
data collected from two genres of bridges in Louisiana: a prestressed concrete slab-on-girder bridge and
a steel bridge with a concrete deck. Although there were limitations concerning the number and type
of sensors used, much information was collected and useful inferences were made that may help better
diagnose the health of bridges monitored in the future using this technique.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
The current state of bridges in the United States calls for the
implementation of a continuous bridge monitoring system that can
aid in timely damage detection and help extend the service life of
these structures. A typical monitoring system would be one which
enables non-invasive, continuous monitoring of the structure. The
passive nature of the acoustic emission (AE) evaluation technique
makes it an ideal choice to serve this purpose. Although the
technique has been successfully used for decades for damage
detection in other fields, its potential in bridge monitoring has not
yet been fully exploited.
Be it for quality control of bridges under construction or structural integrity assessment and monitoring of existing bridges, the
versatility of nondestructive evaluation (NDE) justifies its use in
these structures. Of the many passive NDE techniques available
today, AE was found to be the most widely used for highway
structural assessment [1]. AE testing is a powerful nondestructive
testing tool for real time examination of the behavior of materials
deforming under stress. Load conditions that exist in bridges have
been known to cause materials like concrete and steel to emit energy in the form of elastic waves due to various material-relevant
damage mechanisms. These waves are picked up by sensors
attached to the surface of the material. Further evaluation of the
collected information gives us an overall picture as to the health of
the bridge and helps prioritize repair and maintenance.
This review primarily focuses on the role of AE in bridge monitoring. In the context of bridges, a few merits and limitations of

Corresponding author.
E-mail addresses: anair1@lsu.edu (A. Nair), cscai@lsu.edu (C.S. Cai).

0141-0296/$ see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2010.02.020

the AE technique have also been listed. The first part mainly discusses the basics of AE, covering the topics of equipment requirements and advances, measurement methods, and various available
data processing techniques. This is followed by a brief overview of
the relevant research work completed to date, including both lab
and field tests carried out on concrete, steel, and fiber reinforced
polymer (FRP) bridges or bridge components. Future prospects of
enhancing the capabilities of AE in bridge monitoring systems are
also discussed briefly. Finally, the possibilities of one of the quantitative processing techniques have also been illustrated through
two case studies conducted by the writers.
Although a limited amount of information regarding AE in
bridge monitoring is available, previous reviews presented by authors such as Carter and Holford [2], Holford and Lark [3], and the
ASNT NDT handbook [4] provide quite a comprehensive resource.
Basics of AE pertaining to bridge monitoring are introduced below
for the convenience of the readers.
2. Methodology of acoustic emission
2.1. Basics of acoustic emission
AE is the class of phenomena whereby transient elastic waves
are generated by the rapid release of energy from a localized source
or sources within a material, or the transient elastic wave(s) so generated (ANSI/[5]). Thus, an acoustic monitoring system essentially
requires two integral components: a material deformation that becomes the source, and transducers that receive the stress waves
that are generated from the source.
The schematic shown in Fig. 1 represents the general working principle of an acoustic monitoring system. A developing flaw
emits bursts of energy in the form of high frequency sound waves
which propagate within the material and are received by sensors.

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

Signal

1705

Electronics

One or more Sensors

stimulus
(force)

stimulus
(force)

Fig. 1. Principle of acoustic emission [6].

In general, acoustic emissions can be classified into primary and


secondary emissions. Primary emissions are those originating from
within the material of interest, while secondary emissions refer
to all other emissions generated from external sources. Acoustic
event detection is closely related to the characteristics of stress
waves, such as wave mode, attenuation, effects of multiple paths,
and source location algorithm criterion [3].
Basically, there are two types of AE monitoring strategies that
can be adopted, global and local. A global monitoring helps assess
an entire structures integrity, while local monitoring addresses a
given specific area of damage [2,7]. Based on the duration for which
monitoring is required, AE bridge monitoring may also be categorized into long term and short term. Long term monitoring is
mostly applicable in scenarios where the bridges are relatively new
or require monitoring as a follow up to a regular bridge inspection.
Short term monitoring is more specific, usually the outcome of a
short-term study to update the structures integrity status.
The AE technique has both advantages and disadvantages. The
advantages of the AE technique in the context of bridge monitoring
may be listed as:
1. Damage growth essentially generates AE and is the outcome of
the load history experienced by the structure [8].
2. AE is applicable for local, global, remote, and continuous monitoring purposes without hindering traffic over the bridge structure.
3. Source detection and location algorithms have been improved
to a great extent, assuring reliable analyses.
4. Dynamics of the material are observable in real time due to the
technological advancements made in acquisition systems.
The disadvantages of this method are:
1. Although the issue of background noise discrimination has
been fairly addressed; real-time separation requires several
trial monitoring sessions and experienced personnel.
2. Quantitative AE analyses are still difficult for applications to actual bridge structures.
3. Standardized procedures are not available for all types of
bridges, as most recommendations cater for bridges under
unique conditions of loading, materials, etc.
The measurement of AE essentially involves three basic components: the generated AE wave, the detection equipment used to
capture AE signals, and the processing and interpretation of the
collected data. Understanding the propagation characteristics of
AE waves is vital in distinguishing meaningful data from the unwanted information. AE is an elastic wave that usually has a broadband frequency at the origin. Thus, a typical AE wave detected by
an AE sensor is a combination of longitudinal, transverse, reflected
waves, etc. [9]. The knowledge of wave modes and their characteristics is necessary to implement accurate source location algorithms. The details regarding the theoretical aspect of AE wave
propagation is addressed in numerous works (such as [3], etc.).
Most traditional AE sensors utilized consist of piezoelectric elements undergoing transduction. Their primary function is to detect transient elastic mechanical waves and convert them into

Fig. 2. Common types of AE sensors (pacndt.com).

electrical AE signals. Today, a wide variety of these sensors are


available commercially (Fig. 2). The appropriate transducers
needed for data acquisition are chosen based on the purpose and
sensitivity required for the investigation. Most researchers recommend the use of resonant sensors, as they are highly sensitive to typical AE sources. For bridge monitoring, unidirectional
sensors and sensors sensitive more to in-plane wave modes may
prove beneficial in differentiating AE sources in various bridge
components [2]. The sensors are usually mounted on to the structure, by using adhesives such as epoxy and hot melt glue, or by
holders. Preamplifiers available either separately, or integrated
with the sensor, are integral components that improve the signalto-noise ratio. The signals that are received by the sensors are
collected, stored, and processed in a data acquisition system.
Commercial systems provide customized software that facilitates
qualitative real-time assessment of the collected AE data.
2.2. Interpretation of AE signals
Both the substructure and superstructure of bridges exhibit typical damage modes, which may include corrosion, cracking, and
physical damage due to impact, fire, or fatigue cracking. While
monitoring the superstructure, it must be assessed thoroughly
in critical sections such as shear zones, tension zones, bearing
regions, corrosion prone areas, etc. [10].
Prior to any monitoring procedure, it is essential to understand
all the factors that may influence the AE signature, such as transducer sensitivity, background noise, etc. Special attention must be
paid to the attenuation and wave velocity properties of materials.
Higher frequency sensors tend to exhibit greater attenuation with
distance in steel members [2], while in composite materials such as
concrete and FRP the influence of attenuation is much greater and
plays a crucial role in determining sensor placement and source
location.
Typically, the signals collected can be represented by characteristic parameters such as amplitude, duration, etc., as shown in
Fig. 3. There are numerous qualitative as well as quantitative ways
to interpret these signal parameters or waveforms. For example,
parametric analysis of the AE signal resulted in evaluation criteria
such as: (i) A concrete beam integrity (CBI) ratio, defined as the ratio of the load at the onset of new AE in a subsequent load cycle to
the maximum prior load [11] and (ii) Calm and load ratios of reinforced concrete beams; where the calm ratio is the cumulative
AE activities ratio during the unloading process to the maximum
of the last loading cycle and the load ratio is the ratio between the
load at onset of AE to the prior load [12], etc.
A material under load is known to emit acoustic waves only
after a primary load level is exceeded. This characteristic, named
the Kaiser effect, was first investigated by Joseph Kaiser in 1950.
This effect, portrayed along plot points 1-2-3 in Fig. 4, has been
shown to exist at 70 to 85% of ultimate strength in concrete material. Meanwhile, the Felicity effect (absence of the Kaiser effect) in

1706

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

1000

Rise
Time

Energy

Major
Severity

Volts

Amplitude
Threshold
Time

Follow-up

100

Intermediate
Minor

10

Insignificant
Threshold
Crossing

1
Counts

10
Historic index

Fig. 5. Typical intensity chart for FRP material [18].

Duration

Time

Fig. 3. A typical AE signal [13].

Fig. 4. Kaiser and Felicity effects [16].

composites has led to the use of the Felicity ratio in tracking damage progression in this material [14,15].
Meanwhile, quantification by statistical analysis of parameters
gave rise to the use of Historic and Severity indices in assessing
structural members [17]. This technique has already been successfully applied to FRP and metal piping system evaluations [18]. The
techniques applicability to concrete bridges has previously been
reported by Golaski et al. [17]. A typical intensity analysis evaluates
the structural significance of an AE event by tracking the changes
over time of two indices known as:
(a) The Historic index, which is defined as a measure of the
change in signal strength through the loading phase of the test, and
(b) The Severity index, which is defined as the average signal
strength among the largest numerical values of the signal.
The indices are calculated using the following formulas [19]:

H (I ) =

N
P

Soi

i=K +1
N

N K P
N

(1)

Soi

i=1

Sr =

1
J

J
X

!
Som

(2)

m=1

where, H (I ) Historic index; N number of hits up to time t; Soi


signal strength of the ith hit ; K empirically derived constant

based on material; Sr Severity index; J empirically derived


constant based on material; Som signal strength of the mth hit,
where the order of m is based on the magnitude of the signal
strength.
For concrete, K values are related to N by the relations: N 50,
K = 0; 51 N 200, K = N 30; 201 N 500, K = 0.85N;
and N 501, K = N 75 as well as J values for N < 50, J = 0
and N 50, J = 50 [20,17].
For metals, K values are related to N by the relations: N 15,
K = 0; 16 N 75, K = N 15; 76 N 1000, K = 0.8N;
and N 1001, K = N 200 as well as J values for N < 10, J = 0
and N 10, J = 10 [20].
These indices are evaluated from the signal strength data collected by each sensor. The severity index and maximum value
of historic index is plotted on an intensity chart, which may be
divided into the zones of damage shown in Fig. 5 [18]. The dependence of the technique on the number of data points and
empirically derived constants may be considered limitations of the
technique.
Quantitative analysis of AE signals is mostly carried out with
AE waveforms. Appropriate interpretations of the collected waveforms may be done by subjecting them to any typical waveform
processing tool. b-value analysis of AE signals was yet another
quantitative analysis proposition put forward by Colombo et al.
[21]. Although numerous techniques of AE data assessment have
been proposed and proven useful in laboratory trials, very rarely
have any of them been reported viable for practical onsite monitoring.
Though a standardized procedure is not yet available for monitoring of all bridge structures, the ASTM E1932-02 [22] serves as
a guide for local area short term monitoring of bridge structures.
Recommendations for bridge and/or bridge component monitoring have been compiled by Lozev et al. [23] for steel bridge members, Yuyama et al. [11] for RC beam evaluation, and in 2001 the
Texas Department of Transportation developed the procedure for
monitoring prestressed concrete girders.
3. State-of-the-art of AE bridge monitoring
The primary goal of AE monitoring in structures is to detect,
locate, and assess the intensity of damage [3]. Thus, indicators of
structural damage such as cracks, corrosion, and delamination that
warn of an impending failure have become the focal point of any
AE study. The exceptional features of this monitoring technique are
that there is no issue of traffic interruption during in-service monitoring of the bridges and the use of sensors with small surface areas eliminates any concern about the contact surface profile [1].
Investigations into materials that constitute the civil infrastructure, such as concrete and steel, took place much later compared to other fields where AE is a well-established standard NDE
technique. In the following paragraphs AE monitoring research
conducted on conventional bridge component materials such as
steel and concrete are discussed along with the new generation FRP
bridge components.

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

1707

3.1. Crack monitoring

3.2. Bridge monitoring applications


In addition to the numerous laboratory studies conducted, AE
technology has been used for source location and damage intensity
predictions in numerous field bridge testing applications. Both
short term and long term monitoring instances will be summarized
in the following paragraphs.
One of the pioneering works in AE bridge monitoring was carried out by Pollock and Smith in the 1970s [33]. They monitored a
portable military bridge subjected to proof testing, and reported
amplitude distribution analysis and source location results. The
1980s saw the advent of long-term continuous monitoring implemented on a bridge. The project was spearheaded by the Dunegan
Corporation and was carried out for about 10 months. The purpose
of this study was to check out the feasibility of long-term monitoring. The issue of background noise discrimination led researchers
like Miller et al. [34] to assess both time and frequency domain
AE signals to distinguish various sources. Prine and Hopwood [35]
contributed to this field using an AE weld monitoring system to
locate crack activity in steel bridges. The significance of using
guard sensors to eliminate irrelevant AE events was the outcome
of numerous studies carried out on steel bridges by the Physical
Acoustics Corporation (PAC). Guidelines meant for local and global
monitoring of steel bridges were developed thanks to the contributions of Pollock and Carlyle [36], Carter and Holford [2], Pullin
et al. [37], etc.
Reinforcement corrosion and the resulting cracking are considered the main damage mechanisms that require the need for longterm condition assessment of concrete bridges. Various universities and highway agencies have begun working towards achieving the goal of implementing the AE method for concrete bridges
[38]. Work on concrete structures has been primarily carried
out by Yuyama and Ohtsu [28]. They applied AE to study fracture characteristics, quantifying micro fractures and evaluating
the damage intensity in concrete. Attenuation trends in concrete
bridges were the subject investigated by researchers such as Landis
and Shah [39] and Beck et al. [40]. The NDE Validation Center
(NDEVC) in Virginia conducted acoustic emission tests on concrete
bridges from 19962000. They used the monitoring technique to
detect cracks in bridges by passing high experimental overloads.

1000

E
D

Severity

Numerous laboratory studies have been conducted to demonstrate the ability of AE to detect cracks prematurely. Morton et al.
[24], Holford et al. [25], Hamstad and McColskey [26], etc., have
focused on monitoring fatigue crack development and its correlation with AE activity in steel members. A summary of fracture AE
in metals can be found in [27]. Similarly, cracking in concrete had
been the interest of Yuyama and Ohtsu [28] who primarily used
moment tensor analysis to characterize fracture mechanisms in
RC beams reinforced with reinforcing bars and fiber plastic sheets.
They reported that the breakdown of the Kaiser effect occurred
once shear cracking started to set in, and high AE activity in unloading phases implied serious damage. A detailed study of the
AE waveforms revealed that signals produced as an outcome of
shear and flexural cracking had larger amplitudes and duration
than micro cracking phases of damage [29]. Use of conventional AE
instrumentation has led researchers to come out with conflicting
remarks with regards to trends observed in fiber reinforced polymer (FRP) fracture characteristics. A group of researchers, including Shippen and Adams [30], Gostautas et al. [15], etc. reported
that matrix cracking in FRP produced low amplitude signals, while
Valentin [31], Jamison [32], etc. claimed that matrix cracking was
the higher amplitude source mechanism.

100
C
B

10

A
1

10
Historic index
Fig. 6. Intensity plot for a whole bridge [17].

These kinds of special overloads were used as they could not observe any acoustic activity prior to that point. Drafting of a recommended practice in Poland, for testing reinforced and prestressed
concrete structures by AE, culminated from the research work that
was carried out by Golaski et al. [17]. They reported qualitative results from the testing of five different concrete bridges situated in
Poland at different intensities of damage. Shown in Fig. 6 is the
quantified AE result they plotted from testing a new prestressed
concrete bridge, wherein each dot plotted represents AE signals
characterized by analogous features. Since all points lie in the A
zone, implying no serious deterioration within monitored zones,
the plot aptly portrays the structural health of the new bridge.
In-service RC bridges were also assessed by Beck et al. [40] and
Pullin et al. [41], who reported that the reliability of the AE technique of monitoring bridges was in need of an upgrade.
Bridge cable monitoring using AE had been set up since the
1970s. The successful application of AE in monitoring prestressed
structures [42] inspired Paulson and Cullington [43] to adapt the
technique for continuous monitoring of suspension and cable stay
bridge cables. From several trials, they concluded that AE monitoring is indeed suitable for detecting and locating wire breaks in
cable structures. A similar prestressed concrete bridge application
was reported by Brevet et al. [44]. They observed the effectiveness
of wire fracture monitoring, in a prestressed concrete bridge open
to regular traffic, on cables that cannot be inspected otherwise. AEbased health monitoring approaches for bridge stay cables were
extensively studied by numerous researchers such as Rizzo and Di
Scalea [45], Kretz et al. [46], Fricker and Vogel [47], Li and Ou [48],
Jin et al. [49], etc. Gaillet et al. [50] and Zejli et al. [51] assessed
cable anchorages using AE.
Monitoring of prestressed concrete bridges was also reported
by Vogel et al. [52]. Prestressed concrete structures are known to
have almost no cracks at their initial phase of service life; thus,
Vogel et al. [52] suggested that AE might prove more beneficial in
monitoring new cracks that may develop during their service life.
Owing to the various advantages possessed by Fiber Reinforced
Polymers (FRP) it has become an emerging alternative to the traditional materials that constitute bridge components. Since this
material is still not conventionally used in bridges, the literature
available on AE monitoring of bridge components made of this material is limited. However, the method has been used in various laboratory investigations of FRP bridge decks to study the AE signature
and make valuable correlations. One such study was conducted
by Ziehl and Bane [8], who reported their qualitative approach
to testing a sinusoidal sandwich FRP bridge deck. They devised
a cyclic load profile to enable study of acoustic events generated

1708

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

wireless RF data
transmission

transmitter and
receiver (internet)

Fig. 7. Wireless sensing of bridges using radio frequency transmission [60].

at load holds, and traced damage progression from variations observed in the Felicity ratio. Another successful qualitative assessment was conducted by Kalny et al. [53]. They evaluated the change
in AE signature exhibited by a specimen, before and after repair,
under static loading conditions. They concluded that AE activity
was clearly distinguishable prior to repair, and that pre-existing
damage detection was possible by observing AE activity trends.
Historic and severity indices were the basis on which six full-scale
FRP bridge decks in both original and repaired conditions were
evaluated by Gostautas et al. [15]. Although a clear intensity gradation was not achieved for this unique instance, they reported that
the intensity analysis was useful to identify the onset of damage
and subsequent calculation of the Felicity ratio.
3.3. Advancements in AE equipment technology
Since there has been no one system that has been confirmed as
an ideal bridge monitoring system, various issues with regards to
limited sensitivity of available sensors, practical difficulties faced
during onsite installation, and remote access capabilities have been
investigated over the years. The shortcomings observed while using traditional sensors for structural monitoring have been addressed with proposals for new generation AE sensors that are
much more compact, sensitive, and economically viable. The introduction of micro-electro mechanical systems (MEMS) AE sensors
by Ozevin et al. [54] is one such innovation. These sensors show
promising applications for use in bridge monitoring in conjunction with artificial intelligence networks [55]. Similarly, fiber optics technology is also being explored to develop a new generation
of AE sensors [56].
Obviously, one of the key features desirable in a bridge monitoring situation is remote monitoring. This ability for existing
commercial systems has already been incorporated by corporations, such as PAC, which provide on-line remote web monitoring
facilities. The Local Area Monitoring (LAM) is one such AE monitoring instrument developed by PAC in collaboration with the
Federal Highway Administration (FHWA). Stryk and Pospisisl [57]
proposed developing a monitoring system that identifies rebar corrosion, a crucial concern in concrete bridges. A Canadian company,
Pure Technologies Ltd., has developed SoundPrint, which locates
wire breaks in prestressing tendons [58]. Vallen systems has introduced AMSY4 and AMSY-5 that had continuous sampling rates of
10 MHz for the feature extraction required for real-time data processing [59]. Implementation of wireless AE sensors is yet another
innovation proposed by Grosse et al. [60]. Fig. 7 provides the basic
concept behind remote monitoring intended for AE using wireless technology. Using this technology, along with performanceenhanced sensors based on MEMS, makes this technology more
economic for huge structures such as bridges [60].

16.77m

16.77m

16.77m

Instrumented span

Fig. 8. Sketch of the tested prestressed concrete bridge.

In the following sections, the feasibility of intensity charts


to evaluate the health of bridges is explored by analyzing data
collected from two bridge sites in Louisiana. A prestressed concrete
bridge and a steel bridge under live load conditions were tested.
Both qualitative and quantitative assessments of the AE data
collected from the bridge sites will be presented. Although there
are no standard intensity curves specific to reinforced/prestressed
concrete and steel materials, the general trends observed in an
intensity plot, and the structural assessments that can be made,
are reported here.
4. Case study of a prestressed concrete bridge
This typical prestressed concrete slab-on-girder bridge is located over the Cypress Bayou in District 61 on LA 408 East,
Louisiana. It was built in 1984 and designed to be comprised of
three straight simple spans each 16.77 m (55 ft) in length (Fig. 8).
Each span was supported by 7 AASHTO Type II prestressed concrete girders, with girders spaced 2.13 m (7 ft center to center)
apart. The roadway had a width of 14.33 m (47 ft) and a bridge
deck of 203 mm (8 in) thickness. Each span has one intermediate
diaphragm (ID) located at its mid-span which is not connected to
the deck. Only the third span of the bridge was instrumented.
The bridge was tested for 3 consecutive days in Feb 2006. Both
static and dynamic live load tests were carried out using two
similarly weighing dump trucks. All four of the acoustic sensors
used were R6I (55 kHz resonant frequency) in conjunction with a
DiSP-16 outdoor acquisition work station, both manufactured by
Physical Acoustics Corporation (PAC).
4.1. Loading schedule and equipment setup
One of the main objectives of this test was to assess the
necessity of intermediate diaphragms in prestressed concrete
bridge structures. For this bridge test, acoustic emission monitoring was conducted as a supplementary NDE technique in an attempt only to assess whether there was any additional damage
at the girderdiaphragm connection when subjected to live loads.
Although other gauges such as strain gauges, accelerometers, and

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

(a) SR_T1Sh_P1.

(b) SS_T1L1_P1.

(c) SS_T2L2_P1.

(d) SR_T1L1_T2L2_P1.

1709

Fig. 9. Truck position for various load cases.


Table 1
Load case nomenclature.
Type of live load

Meaning

SR
SS
D30

Static rolling, truck speed < 5 mph


Static stopping, truck mid axle located at mid-span
Dynamic, number following designation
represents the speed of truck (mph)

Truck designation
T1
T2
Roadway designation
L1
L2
Sh
Load case repetition
P1
P2

Truck 1
Truck 2
Lane 1
Lane 2
Shoulder lane

Sensor 4

Sensor 3

Sensor 2

Sensor 1
G7

G5

G6

Fig. 10. Sensor arrangement for test Day 1.

Diaphragm

Pass 1
Pass 2

Sensor 2

Sensor 3

GIRDER #4

deflection gauges were an integral part of this test, the results


pertaining to the acoustic sensors alone will be discussed here.
Since there were limitations on the choice of sensors and their
number, the sensor configurations were chosen based on critical
regions previously predicted from the finite-element model of the
bridge.
The live load tests were carried out with two dump trucks
weighing 271.8 kN (61.1 kips) for Truck 1 and 272.7 kN (61.3 kips)
for Truck 2. Both trucks had a single front axle and two-axles tandem at the rear. The static wheel loads for the first, second, and
third axles of both trucks were about 40.0, 47.8, and 47.8 kN (9,
10.75 and 10.75 kips), respectively.
Since the test consisted of many load cases, a systematic labeling system was developed for identifying the details of each load
case, as shown in Table 1. Thus, a load case named SR_T1Sh_P1 will
translate into a static rolling test case (SR, moving at a speed less
than 5 mph) with truck 1 (T1) over the shoulder lane (Sh) on the
first passage (P1). A few illustrations of load cases accompanied by
their names and cross-sectional details of the monitored span are
shown in Fig. 9.
On the first day of testing, the sensors were placed on the
intermediate diaphragm around girder # 6 (G6 in Fig. 10). The
intention of this configuration was to detect and/or locate any significant changes in stress or the presence of damage at the girder
diaphragm connection region.
The second and third day sensor array consisted of two sensors
(Sensors 1 and 2) being placed 0.61 m (2 ft) apart around the midspan of girder # 4, and the other two were placed on opposite faces
of a section of the intermediate diaphragm close to the same girder.
The chosen sensor arrangements for both days of testing are shown
in Fig. 11.
The data acquisition was carried out with a convenient outdoor
DiSP- workstation unit placed at a location close to the bridge. Realtime monitoring was enabled by the available AEWin software

Sensor1

45.10m
0.61
m

Sensor 4

Fig. 11. Sensor locations for Day 2 and Day 3 (Elevation view).

along with the acquisition system. Prior to acquiring any actual


live load test data, pencil lead break (PLB) tests were carried out
close to all four sensors to ensure their sensitivity. This procedure
also helps in evaluating the attenuation properties in the region
of interest. The suitable sensor spacing is usually determined at a
distance within which the AE amplitude attenuates to about 30 dB
[12]. A suitable threshold level of 45 dB is also chosen at this stage,
based on the background noise level existing at the bridge site.
4.2. Qualitative results
Customized qualitative results are generated by the provided
AEWin software. Only the load cases that generated the most
acoustic activity on each day are discussed in the following section.
The AE activity observed on Day 1, when two trucks were statically moving over the shoulder and lane 1 of the bridge, is indicated
in Fig. 12. The plot reveals that activities recorded by the sensors
located at the lower part of the diaphragm (sensors 1 and 2 in
Fig. 10) generated more acoustic activities relative to the sensors
placed close to the deck. This trend may be attributed to the stress
generated at the monitored joint due to the relative movement between the discontinuous joint at the beamdiaphragm connection
created during loading. Since the observed signal amplitudes were
low, they are not associated with the presence of any live cracks.
Thus, from the amplitude information shown in Fig. 13, one can
infer that the monitored region had no serious structural damage.

1710

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

100
Amplitude (dB)

300
Events

80

240
180

60
40

120
60

50

100

0
40
32

io
n

90

Y position

350

400

100

250

300

350

400

80

Amplitude (dB)

60

sit

po

300

16

250

100

24

30

200

Time (sec)

Fig. 12. Events versus location of sensors for Day 1 configuration (SS_T1Sh_T2L1_P1).

Amplitude (dB)

150

60
40

80

50

100

150

200

Time (sec)

60

Fig. 14. Amplitude versus time plot for load case SR_T1L1_T2L2_P1/SS_T1L1_T2L2_P1
(Top: sensors 1 and 4; bottom: sensors 2 and 3).

40
0

20

40

60

80

100

100

Results for a single load case in the following paragraph are discussed individually for each pair of sensors due to their location
on different parts of the bridge. The Day 2 amplitude plot seen in
Fig. 14 is generated by the two truck load cases, where the first cluster of data points were generated during the trucks rolling phase of
the test (load case SR_T1L1_T2L2_P1) and after a short time lapse
(approximately 220 s), the second cluster of data points were generated after the trucks were backed up and stopped at the midspan
of the bridge (load case SS_T1L1_T2L2_P1). Careful examination of
the amplitude plots in Fig. 14 also reveals the existence of a few
high amplitude events. This may have been caused by secondary
AE sources originating from the structure due to load effects.
Although the sensor configuration on Day 3 was similar to the
previous day (Day 2), on this day the dynamic load case generated the most acoustic activity among all the load cases. This result
could have occurred because of the existence of the Kaiser effect in
concrete mentioned earlier. Since most of the acoustic signal amplitudes lie in the 60 dB range, the activity may not be a result of
any crack-related damage. Relatively higher acoustic activity was
observed at sensors 2 and 3, placed on the diaphragm, than the
other sensors 1 and 4. This could be attributed to the presence of
discontinuities at the girderdiaphragm connection. Upon close visual observation, the bridge girders appear to be in fairly good condition, with virtually crack-free surfaces. This condition is expected
for prestressed concrete bridges. Thus, even though a few high amplitude events were recorded, these may have been contributed to
secondary sources of AE such as relative displacement of the monitored regions due to load effects and concrete-reinforcement interactions at the interface. It was also noted that a better acoustic
response from the bridge was observable when the structure was
subjected to quasi-static loads rather than dynamic loads. A similar observation was also reported by Golaski et al. [17]. A detailed
analysis of the AE waveforms can help to distinguish the various
AE sources (see Fig. 15).

80
60
40
0

10

15

20

Time (sec)
100
Amplitude (dB)

Fig. 13. Amplitude versus time of sensor 2 for load case SS_T1Sh_T2L1_P1.

Amplitude (dB)

Time (sec)

80
60
40
0

10

15

20

Time (sec)
Fig. 15. Amplitude versus time plot for load case D40_T1L2_P1 (Top: sensors 1 and
4; bottom: sensors 2 and 3).

4.3. Damage quantification


To get a better insight into the significance of the AE data collected, quantification of the data is attempted here using the intensity analysis technique. This method requires the accumulation of
AE data obtained from successive load cycles. This AE data is then
used to determine the indices given in Eqs. (1) and (2), which are
summarized in Table 2.
The maximum historic index value and severity indices calculated for each load case shown in Table 2 are plotted on an intensity

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

Intensity chart for Day 1

Intensity Chart for Day2 (Ch# 1 & 4)

10
SR_T1Sh_P1

2
2

Severity

Severity

10

1711

SR_T1L1_T2Sh_P1
SS_T1L1_T2Sh_P1

4
4

4
1
1

SR_T1L1_P1

SR_T1L1_T2Sh_P1

D40_T1L1

10
1

Historic Index

10
Historic index

Fig. 16. Intensity chart for load cases of Day 1 (Numbers within the plot represent
sensor #).

Intensity chart for Day 2 (Ch# 2 & 3)

10

Table 2
Summary of results from intensity analysis.
Ch

H (I )

Sr

Day 1
SR_T1Sh_P1

1
2
4

1.69
4.13
1.79

0.41
2.81
1.18

SR_T1L1_T2Sh_P1

1.72

0.72

SS_T1L1_T2Sh_P1

2
4

4.77
1.80

4.66
2.02

SR_T1L1_P1

Severity

Load case

D40_T1L1

1
1

10
Historic index

Fig. 17. Intensity charts for load cases on Day 2 (Numbers within the plot represent
sensor #).

Day 2

SR_T1L1_P1

1
2
3
4

3.57
1.53
1.93
2.50

1.82
0.13
0.22
1.13

SR_T1L1_T2Sh_P1

1
2
3
4

3.62
2.54
3.34
3.28

2.81
2.11
2.76
1.92

1
2
3
4

3.62
2.08
2.49
3.28

3.31
2.20
2.86
2.23

0.1

SR_T1L1_P1

1.68

0.38

10

SS-T1L1_P1

2
3
4

2.18
2.40
2.14

0.55
0.91
0.50

SR_T1L1_T2L2_P1

1
2
3
4

1.68
2.18
2.40
2.14

0.46
0.73
1.13
0.60

D40_T1L2_P1

1
2
3
4

1.68
2.35
2.40
2.25

0.79
0.95
1.42
1.04

D40_T1L1

SR_T1L1_T2Sh_P1

Intensity Chart for Day 3 (Ch# 1 & 4)

Severity

10

SR_T1L1_P1 SST1L1_P1
1

SR_T1L1_T2L2_P1
1
1

D40_T1L2_P1

Historic index

Severity

Day 3

chart for each day of testing, as shown in Figs. 1618. For the
second and third days of testing, results from sensors 1 & 4, located under the girder, are separated from those of sensors 2
and 3, placed across the thickness of the diaphragm close to the
girderdiaphragm joint. There is no data for some sensors due to
either too small numbers or malfunctions.
The dependence of this analysis technique on a minimum number of data points inhibits the representation of data from every
sensor used for monitoring on the intensity chart. Each intensity
chart has been developed for each day of testing and consecutive
load cases, as the technique requires cumulative data assessment.
Here again, the Day 1 results plotted in Fig. 16 show that sensor 2 seems to have acquired AE events of higher intensities than
all the other sensors. Pre-existing cracks at the girderdiaphragm
interface around the two sides of the observed Girder #6 might
have led to the generation of such acoustic activities. Incremental

10

Intensity chart for Day 3 (Ch# 2 & 3)


SR_T1L1_P1 SST1L1_P1

11

0.1

SR_T1L1_T2L2_P1

3
10

D40_T1L2_P1

Historic index

Fig. 18. Intensity charts for load cases on Day 3 (Numbers within the plot represent
sensor #).

loading leads to consequent intensity points which are plotted towards the right corner of the chart.
The intensity charts plotted for the second and third days
seem to correspond to the qualitative evaluations made previously.
Higher loads generate AE with higher intensities, which in turn
help reflect the intensity of crack-related damage in the monitored
structure. From the plots shown in Figs. 1618, we see that the
points of lower loads plot to the left corner of the chart, while a
higher load causes the intensity point to shift towards the right end
of the chart.
5. Case study of a steel bridge
The bridge that was monitored is located along highway LA1 over the Intracoastal Waterway in Port Allen, Louisiana. The

1712

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

Sensor 3

a
1000

Sensor 4
1.54m Sensor 1

Column

800
Hits

Sensor 2

600

Fig. 19. Acoustic sensor locations on the steel bridge.

400
200
0
0

20

40

60

80

100 120 140 160 180 200

Time (sec)

b
1200
1000
Hits

800
600
400
200

Fig. 20. Oversize load on bridge.

bridge consists of multiple spans with varying span lengths. The


span that was tested is 17.99 m (59) ft long with four steel girders
(W36X182) supporting a concrete deck. The girders were spaced
at 2.64 m (8 ft 8 in center to center). The girders were bolted to
a cross-girder which provides support for them; the cross-girder
was also bolted to the columns. This steel bridge was tested under
overload conditions in July 2006.

0
0

20

40

60

Fig. 21. Cumulative AE hit rate (a) normal traffic phase and (b) overload phase.

100
3

AE monitoring was carried out for two phases of loading: during


normal traffic and overload passage. Since stronger AE activity was
recorded by sensors 3 and 4 located at the beamcolumn joint
for both loading phases, their cumulative hits are represented in
Fig. 21. Upon comparing the signal intensities obtained during both
phases of loading, the following observations may be made:
(a) The signal amplitudes under both loading conditions rarely exceeded 6070 dB. Events in this amplitude range are usually
not associated with any significant structural damage.
(b) The signals obtained from sensors 3 and 4, located at the
beamcolumn joint, were much stronger in comparison with

Ch # 3

3
Sr

5.2. Qualitative results

100 120 140 160 180 200


Time (sec)

5.1. Loading schedule and equipment setup


The objective of this test was to assess the structural behavior
of the monitored span when subjected to overloads. The plan was
to compare the acoustic data collected from normal traffic on the
bridge to that of the overload passage. Since the highway structure
is located near a port, the normal traffic also included heavy trucks.
Potential damage regions such as the mid-span of the girders and
beamcolumn joints were chosen to be monitored. Thus, two of
the sensors were placed under a girder around the mid-span and
the other two were located at the beamcolumn joint, as shown in
Fig. 19.
The optimal threshold level for data acquisition was set at
40 dB. Under the chosen threshold level, no acoustic activity was
observed in the absence of vehicles on the bridge. The oversize
load comprised of two trucks and two trailers, weighing a total of
2401.9 kN (540 kips). The truck was 6.10 m (20 ft) wide and 70.3 m
(230.5 ft) long. The trucks configuration is shown in Fig. 20.

80

Normal traffic
Overlaod

10

Ch # 4
Normal traffic
Overlaod

1
1

H(I)

10

Fig. 22. Intensity chart for acoustic activity from sensors 3 and 4.

the signals recorded by the sensors placed on the girder.


Again, this observation can currently be attributed only as a
source originating from some relative displacement between
the members (such as slip of connection bolts) since no physical damage was observed at that monitored joint after the test.
(c) As expected, the acoustic activity due to the overload is greater
than the normal traffic, which consisted of both light vehicles and heavy trucks. However, the increase (Fig. 21(b)) was
not significant enough to justify more scrutiny in assessing the
monitored region.
5.3. Damage quantification
Intensity charts developed for metal piping systems [15] were
used for the quantitative analysis of data collected from the steel
bridge. Upon analyzing the signal strengths obtained from this test
in both loading conditions, only the data from sensor 3, located on
the beamcolumn joint, during both load conditions could be plotted on the intensity chart. All the other signal intensity values have
a severity value below 1, and thus, were not represented in Fig. 22.
Most of data points fell into the insignificant damage region, except the signal intensities from sensor 3 during overload. Although

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

a signal intensity value lying in zone C implies a defect that requires


a follow-up evaluation, here we cannot assign the high signal intensity to any actual defect, due to various local uncertainties such
as the movement of bolts, proximity to the beamcolumn joint, etc.
in the monitored region.
6. Conclusions and comments
This paper presents a brief review of the research and technology prevalent in acoustic emission monitoring of bridges. Three
materials used in bridge construction: concrete, steel, and FRP have
been considered for this discussion of the research in bridge monitoring. Currently available interpretations of the acquired AE data,
both qualitative and quantitative, have also been discussed. A commendable effort is being made in the direction of improving AE
systems, addressing the practicality and economic issues of implementing the technique for monitoring purposes.
All in all, the applications of AE in bridge monitoring reveal the
potential of this techniques versatility. The technological advances
made in recent years have made the method more suitable for onsite monitoring situations. Although more research may be needed
to implement the current ideas, the future looks promising for the
application of this technology in efficient continuous bridge monitoring scenarios.
The observations and results obtained from the acoustic emission data of two field test cases under live load conditions were also
discussed in this paper. The following conclusions are drawn based
on the observations made through both qualitative and quantitative analyses of the collected AE data.
1. The overall trend seen is that, even with the limited coverage of
the structure, a credible amount of information was collected
and the analysis of this information gives an insight into the
structural response of the local area monitored under live load
conditions.
2. It may also be noted that almost all intensity points plotted on
the data charts for the prestressed concrete bridge had severity
values below 10. This range is considered to represent insignificant emissions in most previously defined intensity charts; and,
thus, leads one to infer that the monitored regions had not developed any significant structural damage during the course of
our testing.
3. The use of intensity charts may help to better estimate the damage severity, although clearly marked zones of damage are not
yet prescribed for certain materials such as concrete and steel.
4. Since the intensity analysis technique assesses cumulative AE
data over successive loads, continued monitoring can help trace
the health of a bridge.
The results obtained from both bridge sites seem to indicate
that the monitored regions had no real issues with their structural
integrity. In spite of the practical challenges faced for use in the
field, continued efforts show that the technique has a promising
future in becoming an integral part of any structural health monitoring system.
Acknowledgements
The authors would like to thank the Louisiana DOTD and
Louisiana Transportation Research Center (LTRC) for making this
study possible. The Louisiana DOTD crew helped conduct the
bridge field tests. Special thanks go to Mr. Walid Alaywan and Artur
DAndreas group. Many graduate students and visiting scholars at
LSU also helped prepare and carry out the bridge test.
References
[1] Rens KL, Wipf TJ, Klaiber FW. Review of nondestructive evaluation techniques
of civil infrastructure. J Perform Constructed Facil, ASCE 1997;11(4):15260.
[2] Carter DC, Holford KM. Strategic consideration for AE monitoring of bridges:
A discussion and Case study. INSIGHT - J British Inst NDT 1998;40(2):1126.

1713

[3] Holford KM, Lark RJ. In: Gongkang Fu, editor. Acoustic Emission testing of
bridges: Inspection and monitoring techniques for bridges and structures.
Cambridge (UK): Woodhead Publishing Ltd; 2005. p. 183215.
[4] American Society of Nondestructive Testing (ASNT). Nondestructive testing
handbook, third edition: vol. 6. Acoustic emission testing, Columbus, OH. 2005
p. 125.
[5] ASTM E1316-07b. Standard terminology for nondestructive examinations.
West Conshohocken (PA): ASTM international; 2007.
[6] Grosse CU. www.NDT.net Editorial: Special issue on acoustic emission. 2002.
[7] Carlos MF, Vahaviolos SJ, Cole PT, Halkyard T. Acoustic emission bridge
inspection/monitoring strategies. In: Structural materials technology IV An
NDT conference. 2000. p. 17983.
[8] Ziehl P, Bane WS. Nondestructive evaluation of fiber reinforced polymer
bridges and decks. FHWA/LA 03/376. LA: Department of Civil and Environmental Engineering, Tulane University. 2003.
[9] Kawamoto S, Williams RS. Acoustic emission and acousto-ultrasonic techniques for wood and wood-based composites: A review. Gen. tech. rep. FPLGTR-134. US Department of Agriculture, Forest Service, Forest Products Laboratory Madison, WI. 2002.
[10] Ghorbanpoor A, Benish N. Non-destructive testing of highway bridges.
Final report # 0092-00-15. Madison (WI): Wisconsin Department of
Transportation; 2003.
[11] Yuyama S, Okamoto T, Shigeishi M, Ohtsu M, Kishi T. A proposed standard
for evaluating structural integrity of reinforced concrete beams by acoustic
emission. In: Vahaviolos SJ, editor. Acoustic emission: Standards and
Technology update, ASTM STP 1353. West Conshohocken (PA): American
Society for testing and materials; 1999.
[12] Ohtsu M, Uchida M, Okamoto T, Yuyama S. Damage assessment of reinforced
concrete beams qualified by acoustic emission. ACI Struct J 2002;99(4):4117.
[13] Huang M, Jiang L, Liaw PK, Brooks CR, Seeley R, Klarstrom DL. Using acoustic
emission in fatigue and fracture materials research. J Mater 1998;50(11).
[14] Ziehl P, Lamanna AJ. Monitoring of the Bonnet carre Spillway bridge during
extreme overload. LTRC project no. 03-6ST. LA: Department of Civil and
Environmental Engineering, Tulane University; 2003.
[15] Gostautas RS, Ramirez G, Peterman RJ, Meggers D. Acoustic emission
monitoring and analysis of glass fiber reinforced composites bridge decks. J
Bridge Eng ASCE 2005;10(6):71321.
[16] Grandt AF. Fundamentals of structural integrity, damage tolerant design and
nondestructive evaluation. NJ: John Wiley and Sons Ltd.; 2003. p. 4268.
[17] Golaski L, Gebski P, Ono K. Diagnostics of reinforced concrete bridges by
acoustic emission. J Acoust Emiss 2002;20:8398.
[18] Committee on Acoustic Emission from Reinforced Plastics (CARP). Recommended practice for acoustic emission of fiberglass reinforced plastic resin
(RP) tanks/vessels. New York: Composites Institute, Society of the Plastics Industry; 1987.
[19] Blessing JA, Fowler TJ, Strauser FE. Intensity analysis. In: Proc., 4th int symp.
on acoustic emission from composite materials. Columbus (Ohio): American
Society for Nondestructive testing; 1992.
[20] Chotickai P. Acoustic emission monitoring of prestressed bridge girders with
premature concrete deterioration. Masters thesis. Austin (Texas): University
of Texas; 2001.
[21] Colombo IS, Main IG, Forde MC. Assessing damage of reinforced concrete beam
using b-value analysis of acoustic emission signals. J Mater Civil Eng, ASCE
2003;15(3):2806.
[22] ASTM E1932-97. e1, Standard guide for acoustic emission examination of small
parts. West Conshohocken (PA): ASTM International; 2002.
[23] Lozev MG, Clemena GG, Duke JC, Sison MF, Horne MR. Acoustic emission
monitoring of steel bridge members. Final report, FHWA/VTRC 97-R13.
Charlottesville (VA): Virginia Transportation Research Council; 1997.
[24] Morton TM, Harrington RM, Bjeletich JG. Acoustic emissions of fatigue crack
growth. Eng Fract Mech 1973;5:6917.
[25] Holford KM, Davies HW, Sammarco A. Analysis of fatigue crack growth in
structural steels by classification by acoustic emission signals. Eng Syst Design
Anal 1994;8:34953.
[26] Hamstad MA, McColskey JD. Detectability of slow crack growth in bridge steels
by acoustic emission. Mater Eval 1999;57(11):116574.
[27] Ono K. New goals for AE in materials research. In: Acoustic emission Beyond
the millennium. UK: Elsevier; 2000. p. 87190.
[28] Yuyama S, Ohtsu M. Acoustic emission evaluation in concrete. In: Kishi T,
Ohtsu M, Yuyama S, editors. Acoustic emission-beyond the millennium.
Elsevier Science Ltd.; 2000. p. 187213.
[29] Yoon DJ, Weiss WJ, Shah SP. Assessing Damage in corroded reinforced concrete
using acoustic emission. J Eng Mech, ASCE 2000;126(3):27383.
[30] Shippen NC, Adams DF. Acoustic emission monitoring of damage progression
in graphite/epoxy laminates. J Reinf Plast Compos 1985;4:24261.
[31] Valentin D. A critical analysis of amplitude histograms obtained during
acoustic emission tests on unidirectional composites with an epoxy and a PSP
matrix. Composites 1985;16(3):22530.
[32] Jamison RD. Microscopic techniques for damage assessment in laminated
composites.
Louthan Jr MR,
LeMay I,
Vander Voort GF, editors.
Microstructural science, vol. 14. American Society for Metals; 1987. p. 53959.
[33] Pollock AA, Smith B. Acoustic emission monitoring of a military bridge.
Nondestr Test 1972;5(6):16486.

1714

A. Nair, C.S. Cai / Engineering Structures 32 (2010) 17041714

[34] Miller RK, Ringermacher HI, Williams RS, Zwicke PE. Characterization
of acoustic emission signals. Report no. R83-996043-2. East Hartford
(Connecticut): United Technologies Research Center; 1983.
[35] Prine DW, Hopwood T. Detection of fatigue cracks in highway bridges with
acoustic emission. J Acoust Emiss 1985;4(23):S3046.
[36] Pollock AA, Carlyle JM. Acoustic emission for bridge inspection: Application
guidelines. Final report, Contract DTFH61-90-C-0049. Washington (DC):
Federal Highway Administration; 1995.
[37] Pullin R, Carter DC, Holford KM. Damage assessment in steel bridges. In: Key
engineering materials. Switzerland: Trans Tech Publications; 1999. p. 1678;
33542.
[38] Watson JR, Yuyama S, Pullin R, Ing M. Acoustic emission monitoring
applications for civil structures, Bridge Management, Thomas Telford, London,
2005. p. 56370.
[39] Landis EN, Shah SP. Frequency-dependent stress wave attenuation in cementbased materials. J Eng Mech 1995;121(6):73743.
[40] Beck P, Bradshaw TP, Lark RJ, Holford KM. A quantitative study of the
relationship between concrete crack parameters and acoustic emission energy
released during failure. In: Key engineering materials. Switzerland: Trans Tech
Publications; 2003. p. 2456; 4616.
[41] Pullin R, Holford KM, Lark RJ, Beck P. Acoustic emission assessment of
concrete hinge joints. In: Key engineering materials. Switzerland: Trans Tech
Publications; 2003. p. 2456; 32330.
[42] Elliot JF. Monitoring of prestressed structures. Civil Eng, ASCE 1996;66(7):
613.
[43] Paulson PO, Cullington DW. Evaluation of continuous acoustic monitoring as
means of detecting failures in posttensioned and suspension bridges. In: XIII
FIP congress & exhibition. 1998.
[44] Brevet P, Robert JL, Aubaagnac A. Acoustic emission monitoring of Bridge
cables: Application to a Pre-stressed Concrete bridge. DEStech Publications,
First European Workshop on Structural Health Monitoring, SHM 2002, ENS
Cachan France. 2002. p. 28793.
[45] Rizzo P, Di Scalea FL. Acoustic emission monitoring of CFRP cables for cablestayed bridges. In: Proceedings of SPIE The International Society for Optical
Engineering, vol. 4337. 2001. p. 12938.
[46] Kretz T, Brevet P, Cremona C, Godart B, Paillusseau P. Continuous monitoring
and structural assessment of the Aquitaine suspension bridge. Bull LPC 2006;
1332.
[47] Fricker S, Vogel T. Site installation and testing of a continuous acoustic
monitoring. Constr Build Mater 2007;21(3):50110.

[48] Li D, Ou J. Acoustic emission monitoring and critical failure identification


of bridge cable damage. In: Nondestructive characterization for composite
materials, aerospace engineering, civil infrastructure, and homeland security
2008, Proceedings of SPIE - The International Society for Optical Engineering,
vol. 6934. 2008. p. 15.
[49] Jin T, Sun Z, Sun L. Acoustic emission monitoring of stayed cables based
on wavelet analysis. In: Sensors and smart structures technologies for
civil, mechanical, and aerospace systems 2008, Proceedings of SPIE The
International Society for Optical Engineering, vol. 6932. 2008. 1-7.
[50] Gaillet L, Tessier C, Bruhat D, Michel R. Diagnostic assessment of multi-layer
cable anchorages by means of acoustic emissions. Bull LPC 2004;2501; 5563.
[51] Zejli H, Laksimi A, Tessier C, Gaillet L, Benmedakhen S. Detection of the broken
wires in the cables hidden parts (anchorings) by acoustic emission. In: Acoustic
emission testing Proceedings of the 27th European conference on acoustic
emission testing. Advanced materials research. 2006. p. 134; 34550.
[52] Vogel T, Schechinger B, Fricker S. Acoustic emission analysis as a monitoring
method for prestressed concrete structures. In: Proceedings EC NDT 9th
European conference on NDT. 2006.
[53] Kalny O, Peterman RJ, Ramirez G. Performance evaluation of repair technique
for damaged fiber-reinforced polymer honeycomb bridge deck panels. J Bridge
Eng ASCE 2004;9(1):7586.
[54] Ozevin D, Greve DW, Oppenheim IJ, Pessiki S. Steel plate coupled behaviour of
MEMS transducers developed for acoustic emission testing. In: 26th European
conference of acoustic emission testing. 2004. p. 55764.
[55] Hay TR, Hay DR, Hay JR, Greve DW, Oppenheim IJ. Transforming bridge
monitoring from time-based to predictive maintenance using acoustic
emission MEMS sensors and artificial intelligence. In: 7th world congress on
railway research. 2006.
[56] Spillman Jr WB, Claus RO. Optical-fiber sensors for the detection of acoustic
emission. MRS Bull 2002;3969.
[57] Stryk J, Pospisisl K. Rebar corrosion in concrete bridges and its detection by
acoustic emission method. CDV- Transport research center, Czech Ministry of
Transportation and Communications. 2001. p. 16.
r
[58] Paulson PO, Elliott JF, Youdan DG. SoundPrint
acoustic monitoring to confirm
integrity of stressed wire in bridges, structures and water pipelines. In: 15th
world conference on nondestructive testing. 2004.
[59] Vallen DIH. AE testing fundamentals, equipment, applications. NDT Net 2002;
7(9).
[60] Grosse CU, Finck F, Kurz JH, Reinhardt HW. Monitoring techniques based on
wireless AE sensors for large structures in civil engineering. In: Proc. EWGAE
2004 symposium. 2004. p. 84356.

You might also like