Effects of Elevated Temperature on Guided-Wave

download Effects of Elevated Temperature on Guided-Wave

of 17

Transcript of Effects of Elevated Temperature on Guided-Wave

  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    1/17

    http://jim.sagepub.com/

    StructuresJournal of Intelligent Material Systems and

    http://jim.sagepub.com/content/19/12/1383The online version of this article can be found at:

    DOI: 10.1177/1045389X07086691

    2008 19: 1383 originally published online 20 May 2008Journal of Intelligent Material Systems and StructuresAjay Raghavan and Carlos E.S. Cesnik

    Effects of Elevated Temperature on Guided-wave Structural Health Monitoring

    Published by:

    http://www.sagepublications.com

    can be found at:Journal of Intelligent Material Systems and StructuresAdditional services and information for

    http://jim.sagepub.com/cgi/alertsEmail Alerts:

    http://jim.sagepub.com/subscriptionsSubscriptions:

    http://www.sagepub.com/journalsReprints.navReprints:

    http://www.sagepub.com/journalsPermissions.navPermissions:

    http://jim.sagepub.com/content/19/12/1383.refs.htmlCitations:

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/content/19/12/1383http://jim.sagepub.com/content/19/12/1383http://www.sagepublications.com/http://jim.sagepub.com/cgi/alertshttp://jim.sagepub.com/cgi/alertshttp://jim.sagepub.com/subscriptionshttp://jim.sagepub.com/subscriptionshttp://jim.sagepub.com/subscriptionshttp://www.sagepub.com/journalsReprints.navhttp://www.sagepub.com/journalsReprints.navhttp://www.sagepub.com/journalsPermissions.navhttp://jim.sagepub.com/content/19/12/1383.refs.htmlhttp://jim.sagepub.com/content/19/12/1383.refs.htmlhttp://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/content/19/12/1383.refs.htmlhttp://www.sagepub.com/journalsPermissions.navhttp://www.sagepub.com/journalsReprints.navhttp://jim.sagepub.com/subscriptionshttp://jim.sagepub.com/cgi/alertshttp://www.sagepublications.com/http://jim.sagepub.com/content/19/12/1383http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    2/17

    Effects of Elevated Temperature on Guided-waveStructural Health Monitoring

    AJAY RAGHAVAN AND CARLOS E. S. CESNIK*

    Department of Aerospace Engineering, The University of Michigan, 1320 Beal AvenueAnn Arbor, Michigan 48109, USA

    ABSTRACT: Elevated temperatures can cause significant changes in guided-wave (GW)propagation and transduction for structural health monitoring (SHM). This work focuses onGW SHM using surface-bonded piezoelectric wafer transducers in metallic plates for thetemperature range encountered in internal spacecraft structures (201508C). First, studiesdone to determine a suitable bonding agent are documented. This is then used in controlledexperiments to examine changes in GW propagation and transduction using PZT-5Apiezoelectric wafers under quasi-statically varying temperature (also from 20 to 1508C).Modeling efforts to explain the experimentally observed increase in time-of-flight and changein sensor response peak-to-peak magnitude with increasing temperature are detailed.Finally, these results are used in detection and location of mild and moderate damage usingthe pulse-echo GW testing approach within the temperature range.

    Key Words: structural health monitoring, damage prognosis, guided waves, Lamb waves,

    thermal variation, temperature compensation, spacecraft structures, integrated systems health

    management, bonding agent.

    INTRODUCTION

    Motivation

    IN 2005, a new vision was defined for NASA, and it

    has already set in motion plans for returning

    astronauts to the Moon, and eventually, longer-term

    missions to Mars. In this endeavor, integrated systemshealth management (ISHM) will play a key role in

    fulfilling the mission objectives. ISHM will help in

    transitioning from low-earth orbit missions with con-

    tinuous ground support to more autonomous long-term

    missions (Carlos et al., 2005). The ISHM system will

    manage all the critical spacecraft functions and systems.

    It will include monitoring capability composed of

    sensors and actuators plus a reasoning system that can

    evaluate the hardwares functionality. The structural

    health monitoring (SHM) subsystem will be a crucial

    component of the ISHM system. It will apprize

    astronauts on changes in vehicle structural integrity

    requiring action as well as providing the crew with thecapability to forecast potential problems and schedule

    repairs based on the rate of loss of system function and

    condition of hardware.

    Fundamentals of Guided-wave Approaches

    Among various technologies under investigation for

    SHM, there are guided-wave (GW)-based approaches

    These essentially involve exciting the structure with

    high frequency GWs and processing the difference in

    structural response with respect to a baseline signal for

    the pristine condition using a tested algorithm to detectdamage and characterize it, if present. Guided waves can

    be defined as stress waves forced to follow a path

    defined by the material boundaries of the structure

    For example, when a beam is excited at high frequency

    stress waves travel in the beam along its axis away from

    the excitation source, i.e., the beam guides the waves

    along its axis. While several transducers have been

    tested, piezoelectric wafer transducers (hereafter referred

    to as piezos) seem to be the most commonly used

    option. This is largely because of the low mass and space

    penalty associated with incorporating them (crucial in

    aerospace structures) and their high energy density for

    high frequency applications. There are two mainapproaches commonly used in GW SHM, pulse-echo

    and pitch-catch. In the former, typically after exciting

    the structure with a pulse, a sensor (usually immediately

    adjacent to the actuator) is used to detect scattered

    echoes of the pulse coming from discontinuities. Since

    the boundaries and the wavespeed for a given cente

    actuation frequency of the toneburst are known*Author to whom correspondence should be addressed.E-mail: [email protected] 16 and 823 appear in color online: http://jim.sagepub.com

    JOURNAL OF INTELLIGENT MATERIAL SYSTEMS AND STRUCTURES, Vol. 19December 2008 1383

    1045-389X/08/12 138316 $10.00/0 DOI: 10.1177/1045389X07086691 SAGE Publications 2008

    Los Angeles, London, New Delhi and Singapore

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    3/17

    the signals from the boundaries can be filtered out

    (assuming a narrow enough time-span excitation pulse is

    used; alternatively one could subtract the test signal

    from the baseline signal). One is then left with signals

    generated from the damage sites (if present). From these

    signals, damage sites can be located using the known

    wavespeed. One potential caveat of this approach isthat there are small areas in the immediate vicinity of

    the actuator and structural features (e.g., stiffeners) with

    decreased sensitivity for the actuatorsensor pair. The

    actual size of these areas decreases with decreasing

    time-span of the excitation pulse. This results from the

    difficulty in distinguishing between minor differences in

    the actuation signal/reflections from structural features

    (e.g., caused by small temperature changes) and reflec-

    tions caused by structural damage.

    In the pitch-catch approach, a pulse signal is usually

    sent across the specimen under interrogation and a

    sensor at the other end of the specimen receives the

    signal (in plate-like structures this is often implementedusing a network of such paths to cover the structure).

    From various characteristics of the received signal, such

    as delay in time-of-flight, amplitude, frequency content,

    etc., information about the damage (if present) can be

    obtained. Thus, the pitch-catch approach cannot be

    used to locate the defect unless a dense network of

    transducers is used. Issues with both approaches in the

    face of changing temperature will be discussed later in

    this study. A detailed survey of GW SHM, including

    fundamentals and early history, is presented in a review

    paper by the authors (Raghavan and Cesnik, 2007c). As

    pointed out there, while GW SHM has shown a good

    deal of promise in various laboratory demonstrations,several issues remain to be resolved before they see

    widespread field deployment in structures. Among these,

    compensation for environmental and service conditions

    is a crucial one.

    Spacecraft Structures and Their Environment

    Typical spacecraft structures are composed of different

    substructures, each of which can act as waveguides,

    thereby making them attractive application areas for GW

    SHM. This is also true of the planned NASAs crew

    exploration vehicle (CEV). The CEV is expected to have

    an aluminum alloy internal structure in the shape ofa blunt body capsule protected by bulk insulation,

    composite skin panels, and a thermal protection system

    (TPS), see Stanley et al. (2005). Spacecraft structures

    in particular present a challenging application due to

    the harsh environment of outer space as well as the

    tremendous heat flux and high temperatures attained

    during re-entry into a planets atmosphere. While fiber

    optic sensors can be designed to withstand these harsh

    re-entry temperatures, they are passive and cannot be

    used to excite GWs. Some authors have examined SHM

    in spacecraft structures using fiber optic sensor networks

    by passive strain/loads monitoring (e.g., Friebele et al.,

    1999; Ecke et al., 2001). However, such passive

    approaches require a much denser network of sensors

    than GW approaches. The internal spacecraft structures,

    however, are somewhat insulated by the TPS. The TPS is

    typically designed to keep temperatures below 1508C ininternal structures, particularly in manned missions

    (Myers et al., 2000). Apart from the re-entry phase,

    even in the course of the flight, the temperature of

    spacecraft structures varies significantly, with tempera-

    tures up to 708C (Larson and Wertz, 1995) depending on

    whether they face towards or away from the Sun. For

    solar arrays, this fluctuation is even greater (up to 1008C,

    see Larson and Wertz, 1995). Another source of

    temperature variation in internal spacecraft structures is

    the heat radiated by cabin electronics, which is difficult to

    reject into space, and is therefore controlled by active

    cooling. Commercial piezos are functional without loss in

    properties up to half their Curie temperature. For leadzirconium titanate, or PZT 5A (a.k.a. DoD Type II), one

    of the more commonly used piezoceramics, half the Curie

    limit is about 1758C. Thus, internal spacecraft structures

    become a potential application area for GW SHM using

    PZT-5A piezos. Some studies have examined GW SHM

    for cryogenic tanks (Blaise and Chang, 2001; Lin et al.,

    2003) and thermal protection systems (Yang et al., 2003;

    Derriso et al., 2004; Huang et al., 2005) in spacecraft

    structures. However, the experiments/simulations in

    these studies were restricted to room temperature

    (except in Blaise and Chang, 2001). The GW SHM

    algorithm must account for temperature changes to

    minimize false damage indications and reduce errors indamage characterization during the course of space

    missions. The present study explores this very issue.

    Previous Efforts on Effects of Temperature for GW SHM

    There have been some efforts to address the issue of

    varying temperature for GW SHM in the literature.

    Blaise and Chang (2001) investigated the performance

    of piezoelectric transducers (in GW pitch-catch config-

    uration) embedded into sandwich structures for cryo-

    genic fuel tanks at low temperatures (up to 908C).

    An empirical model (linear) was fitted to experimentally

    obtained data points for changing signal peak-to-peakmagnitude and time-of-flight. Reasonable agreement

    between the interpolated signals from the empirical

    model and experimental data for intermediate tempera-

    ture values was obtained. However, no damage detec-

    tion studies were reported by them. Lee et al. (2003)

    studied the effect of temperature variation on the Lamb-

    wave response of a piezoceramic sensor in a pitch-catch

    configuration on a metallic plate from room tempera-

    ture up to 708C. They observed that the effect of

    temperature variation over this range (analyzed using

    1384 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    4/17

    principal component analysis) was much more

    pronounced than the effect of damage (drilled 1-mm

    diameter hole). Lu and Michaels (2005) and

    Konstantinidis et al. (2006) examined GW SHM under

    mild thermal variations from 20 to 408C. Both addressed

    modeling the varying time-of-flight in this temperature

    range due to changing substrate elastic modulus andthermal expansion, and good agreement with experi-

    ments in this temperature range was observed. Lu and

    Michaels (2005) also suggested using a bank of baseline

    signals for various temperatures and picking a baseline

    signal which minimizes difference relative to the test

    signal for that particular temperature. Chambers et al.

    (2006) suggested test procedures for environmental

    robustness certification of GW SHM transducers in

    aircraft structures. Schulz et al. (2003) studied the

    performance of PZT-5A patches as free vibration

    sensors bonded using various adhesives on aluminum

    beams up to 2408C. A drop of 74% in strain response

    amplitude (relative to room temperature value) wasobserved at 1508C, and the response dropped to zero at

    2408C. They also explored alternative materials for high

    temperature applications, and identified lithium niobate

    and nanotubes as potential transducer materials for

    future high temperature applications (up to 10008C).

    However, these have much weaker piezoelectric proper-

    ties at room temperatures compared to PZT. Results

    from preliminary thermal experiments were also

    reported by the authors (Raghavan and Cesnik,

    2005a). These tests examined sensor response variation

    in pitch-catch tests done using PZT-5A piezos bonded

    on a 1-mm thick aluminum plate up to 1508C. It was

    observed that the sensor response decreased continu-ously with increasing temperature and went to the noise

    floor beyond 1108C. It was suspected that this decrease

    was due to bonding agent degradation.

    Objectives of this Study

    It is evident from the literature reviewed that the

    issues of compensation for and damage characterization

    under thermal variations expected in GW SHM for

    spacecraft structures (above room temperature) have

    not received much attention. This study aims to

    contribute in these aspects. First, studies done to find

    a suitable bonding agent (for GW SHM using piezo-ceramics on aluminum plates) that does not degrade

    under thermal variations from 20 to 1508C are reported.

    With a suitable bonding agent chosen, controlled

    experiments are done to examine changes in GW

    propagation and transduction using PZT-5A piezos

    under quasi-statically varying temperature (also from

    20 to 1508C). All parameters changing with temperature

    are identified and quantified based on data from

    the literature. Modeling efforts exploiting these data

    to explain the experimental results are outlined.

    Finally, these results are used to explore detection

    and location of damage (indentations/holes) using the

    pulse-echo GW testing approach in the same tempera

    ture range.

    BONDING AGENT SELECTION

    Bonding Agents Tested and Specimen Preparation

    After an initial pre-screening, three different two-par

    epoxies were evaluated for the temperature range o

    interest. These were 10-3004 (Epoxies, Etc., 2006), and

    Epotek 301 and 353ND (Epoxy Technology, 2006)

    Epotek 301 and 353ND, both low-viscosity agents, are

    rated for continuous operation up to 200 and 2508C

    respectively. 10-3004 is relatively viscous, and is rated

    for continuous operation up to 1258C, although the

    manufacturer clarified that it should work up to 1508C

    for short-term use (hours). In addition, it was confirmedfrom the manufacturers that each epoxy would be

    suitable for surface-bonding piezoceramics (with metal-

    lic electrodes) on aluminum plates. While 10-3004 and

    301 can be cured overnight at room temperature

    353ND needs to be cured in an oven at 808C fo

    25 min. Standard surface preparation procedures were

    followed with each, i.e., the plate surface was made

    rough by light sanding, and both the plate and piezos

    were cleaned thoroughly using acetone to get rid of

    grease and dust. After uniformly applying a thin-layer

    film of epoxy to both surfaces and cleaning the excess

    light pressure was applied using small weights (2 lb.) to

    the interface to help the bond set.

    Experimental Tests with Epotek 301 and

    Epoxies, Etc. 10-3004

    The first aluminum alloy (5005) plate specimen

    (40350.32cm3) tested had four PZT-5A piezo

    that were surface-bonded using Epotek 301 (Figure 1)

    Two piezos were used as actuators (dimensions

    2.51.50.03 cm3, at the center, on either surface)

    and two as sensors (dimensions 1 10.03 cm3). One o

    the sensors (Sensor 1) was immediately adjacent to the

    actuator, and the other (Sensor 2) was 10 cm away from

    the plate center. This specimen was thermally cycledfrom 20 to 1508C in an industrial oven and then cooled

    back to room temperature over three cycles. A Labview

    based automated thermal test setup was developed for

    these experiments. After turning the oven on, at every

    108C intervals (read by a type-K thermocouple with

    18C accuracy attached to one side of the plate

    specimen), the Labview program triggered an Agilen

    33220A function generator to send a 3.5-cycle Hanning-

    windowed toneburst, with center frequency 210 kHz to

    the actuators (excited symmetrically), 16 times each

    Effects of Elevated Temperature on Guided-wave SHM 1385

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    5/17

    at 1-s intervals. A Hewlett Packard 54831B Infiniium

    oscilloscope recorded the sensor response signals, which

    were sampled at 10 MHz and averaged over the 16

    readings at each temperature. In these tests, it was

    observed that the sensor response signal of Sensor 2decreased monotonically in peak-to-peak magnitude

    with increasing temperature (Figure 2). The error bars

    shown are based on the standard deviation over the 16

    readings at each temperature. Furthermore, Sensor 2s

    response signal peak-to-peak magnitude at room tem-

    perature decreased at the end of each cycle, and the

    shape of the signal also changed significantly, as shown

    in Figure 3. It should be noted that the sensor response

    was compensated for varying actuation signal magni-

    tude (which dropped due to the increasing capacitance

    of the actuators with temperature). While some amountof irrecoverable loss in response strength is expected

    after the first few cycles, due to thermal pre-stabilization

    of piezos (Berlincourt et al., 2007), the signal shape is

    not expected to change. Despite the actuators being

    excited symmetrically, and thereby only supposedly

    exciting the S0 mode, experimental imperfections cause

    weak A0 mode excitation. To counter this, the sensor

    was originally designed (Raghavan and Cesnik, 2005b)

    to be very weakly sensitive to the A0 mode (the sensor

    size equaled the A0 mode wavelength at 210 kHz).

    However, after each cycle, the strength of the slower A0mode contribution at 208C in the Sensor 2 signal

    increased. This suggested that the sensors effectivearea kept decreasing after each cycle. Based on these

    factors, it was concluded that the Epotek 301 bond line

    was indeed degrading as a result of the thermal cycling.

    An analogous test was done with PZT-5A transducers

    bonded using 10-3004 on an aluminum alloy (5005)

    plate. In this case, the results were even more drastic and

    the sensor response dropped gradually to the noise floor

    at 1008C while heating in the very first cycle, and never

    recovered.

    Experimental Tests with Epotek 353ND

    Finally, tests were done with Epotek 353ND.The specimen tested (also an aluminum 5005 specimen

    of size 4035 0.32 cm3) was similar to the ones tested

    above. The schematic of this is shown in Figure 4. It had

    two 2 10.03 cm3 piezos surface-bonded on either face

    of the plate at the center which were used as actuators.

    Three surface-bonded 1 10.03cm3 piezos were used

    as sensors, of which one was immediately adjacent to one

    of the actuators and two were at a distance of 10.2 cm

    from the plate center. The specimen was thermally cycled

    in the same temperature range seven times in the oven,

    35cm PZT-5A actuators2.5cm x 1.5cm

    10 cm

    3.2 mm

    PZT-5ASensor 2

    PZT-5ASensor 1

    1 cm x 1 cm1 cm x 1 cm

    Aluminum 5005 plate

    Type K thermocouple

    0.3 mm

    40 cm

    Supportblocks

    Figure 1. Schematic of specimen for tests with Epotek 301.

    2 4 6 8 10

    x105

    0.12

    0.1

    0.08

    0.06

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    Time (s)

    Sensor

    2signal(V)

    Before thermal expts.

    After thermal cycle 1

    After thermal cycle 2

    After thermal cycle 3

    EMI

    S0mode A0mode

    Boundaryreflections

    Figure 3. Sensor 2 signal at room temperature before and aftereach of the three thermal cycles for tests with Epotek 301 (thesignals are offset by a small DC voltage for clarity; EMI electromag-netic interference from the actuation).

    0

    0.05

    0.1

    0.15

    0.2

    0 20 40 60 80 100 120 140

    Sensor

    2peak-to-peakmagnitude(V) 1st cycle heating 1st cycle cooling

    2nd cycle heating 2nd cycle cooling

    3rd cycle heating 3rd cycle cooling

    Temperature (deg C)

    Noise floor

    Figure 2. Variation of Sensor 2 response magnitude (peak-to-peak) and associated error bars with temperature over three thermalcycles (for tests with Epotek 301).

    1386 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    6/17

    using the same experimental setup and settings described

    above. In this case, the sensor response peak-to-peak

    magnitude and shapes did not change (within negligible

    error margins, see Figure 5) when the signals before and

    after each thermal cycle are compared. The very firstcycle was an exception, being the thermal pre-stabiliza-

    tion cycle discussed before, which caused a 17% drop in

    Sensors 2 and 3 response peak-to-peak magnitude (but

    the signal shape did not change after the cycle). Thus, this

    epoxy proved to be suitable for the purposes of this study.

    Thereafter, more controlled tests were conducted with

    this same specimen to study signal changes for pristine

    specimens and to explore damage characterization

    at different temperatures, which is discussed in the

    following sections.

    EFFECTS OF ELEVATED TEMPERATURE

    FOR PRISTINE SPECIMEN: EXPERIMENTS

    AND MODELING

    Experimental Setup and Results

    In designing the specimen for tests with Epotek

    353ND, Sensor 1 was collocated with the actuator

    with the intention of using it for damage detection using

    pulse-echo tests. Sensors 2 and 3 were for tracking

    changes in the GW transmitted signal with temperature

    (in undamaged state). In addition, Sensors 2 and 3

    also act as mild GW scatterers, due to the increased

    local stiffness and mass caused by their presence

    This simulates the effect of some structural feature

    (e.g., rivets) which could act as GW scatterers in morecomplex structures. While the specimen was tested in the

    industrial oven initially to check for bond degradation

    the ovens heating/cooling rate could not be tightly

    controlled, and was very rapid (up to 108C/min) a

    times. This fast heating rate led to non-repeatable

    signals for Sensor 1, which could potentially be

    interpreted as false positives. This is discussed in the

    next section. More controlled tests were subsequently

    done in a computer-controlled autoclave (Figure 6)

    where both the heating and cooling rates were set to

    18C/min. A 5-min dwell period at 1508C was also

    included in the thermal cycle between the heating and

    cooling phases (Figure 7). The data at 90 and 1008Cwhile cooling was not used, since in this temperature

    range, the autoclave switches from exclusively air

    cooling to a combination of air and water cooling

    leading to oscillations in the cooling rate over this range

    For these tests, the center frequency was reduced to

    120 kHz. This was to minimize actuation signal distor

    tion effects at higher frequencies caused by increasing

    actuator capacitance at higher temperatures. While

    a Krohn-Hite 7500 wideband amplifier was tried for a

    couple of thermal cycles in the oven, it was unable to

    35cm

    Piezo actuators2 cm x 1 cm10.2

    cm

    3.2 mm

    Aluminum 5005 plate

    Sensor 31 cm x 1 cm

    Thermocouple

    Damage location

    8 cm

    0.3 mm

    Sensor 2

    1 cm x 1 cm

    40 cm Supportblocks

    1 cm x 1 cmSensor1

    Figure 4. Schematic of specimen for tests with Epotek 353ND.(Damage introduced and discussed later.).

    0 1

    x104

    0.05

    0.04

    0.03

    0.02

    0.01

    0

    0.01

    0.02

    0.03

    0.04

    0.05

    Sensor2response(V)

    Before thermal cycleAfter thermal cycle

    S0 mode

    Boundary reflections

    Mild A0mode

    Electromagneticinterference

    Time (s)

    0.2 0.4 0.6 0.8

    Figure 5. GW signal sensed by Sensor 2 (bonded using Epotek3 53 ND ) bef or e a nd a fte r a th er m al c yc l e: th e dif f ic ul ty in distinguishing between the two signals is indicative of therepeatability of signals after the thermal cycle.

    Autoclave

    Plate specimenand cable stand

    TC

    Oscilloscope

    Labview system

    Function generator

    Data acquisition for TC

    Figure 6. Labeled photograph of setup and autoclave for controlledthermal experiments (TC thermocouple).

    Effects of Elevated Temperature on Guided-wave SHM 1387

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    7/17

    amplify without significant signal distortion and ripple

    at higher temperatures. Therefore, no amplifier was usedfor the controlled tests in the autoclave. The actuation

    signal was still a 3.5-cycle Hanning windowed toneburst

    of 18 V peak-to-peak magnitude at 208C, with the two

    actuators on either surface excited symmetrically. At

    every 108C intervals, the LABVIEW setup triggered the

    function generator to send this actuation signal to the

    function generator 30 times with 1-s gaps between each

    trigger. The 1-s gap between each excitation ensures that

    the GWs from the previous excitation have died out and

    do not interfere with the GWs from the next excitation

    burst. The oscilloscope in turn is triggered by a square

    wave pulse from the function generator through a

    separate synchronization channel coinciding with thebeginning of each excitation burst. This ensures that

    the start of the actuation pulse is set as time zero for

    the signal in the oscilloscope. The oscilloscope stores the

    signals from the piezos, averaged over the 30 signals

    from the multiple excitation bursts to reduce signal noise

    levels, along with the standard deviation in peak-to-peak

    magnitude. During this 30-s period over which signal

    averaging is done, the autoclave continues heating/

    cooling at 18C/min, meaning the specimen temperature

    varies by up to 0.58C. However, this temperature

    variation is within the accuracy of the thermocouple

    (18C). It is also representative of situations encoun-

    tered in practical applications: one can seldom expectthe structure to remain at perfect thermal equilibrium

    for on-the-field applications. Data were collected over

    two thermal cycles for the pristine, undamaged condi-

    tion. As mentioned before, as temperature increases, due

    to increasing actuator capacitance, there is an increase in

    the electrical load seen by the function generator.

    Therefore, due to the lack of an amplifier, there was a

    drop in actuation peak-to-peak magnitude from 18 V at

    208C to %13V at 1508C (but negligible shape distor-

    tion). All signals presented for higher temperatures in

    this study have been scaled for 18-V peak-to-peakactuation level (by multiplying the sensor signals with

    the ratio of 18V to the actuation peak-to-peak

    magnitude at the corresponding temperature).

    Figure 8 shows the GW signal read by Sensor 2 at

    various temperatures while heating. Evidently, there is a

    decrease in GW speed of the first transmitted GW pulse

    as temperature increases. In addition, the signal peak-

    to-peak magnitude increases with increasing tempera-

    ture up to a certain point (around 908C) and then

    decreases with increasing temperature. Hysteresis effects

    were found to be negligible (unlike in the oven tests,

    where significant hysteresis was observed between the

    heating and cooling phases due to very differenttemperature change rates in the two phases).

    Identification and Quantification of Thermally

    Sensitive Parameters

    In order to explain these effects, an effort was made to

    identify all parameters in the experiment that change

    with temperature. The following list was compiled and

    data for their thermal variation were found from various

    sources in the literature:

    1. Youngs moduli of structural substrate and PZT-5A:

    The substrate elastic modulus is probably the mostimportant parameter for thermal variations. There is

    a significant decrease in the elastic modulus of

    aluminum with increasing temperature. This causes

    a reduction in GW speeds, as reflected in the change

    in dispersion curves. Furthermore, in quantifying

    thermal variations of elastic modulus, two different

    data sets were found: one for the variation in static

    elastic modulus (U.S. Munitions Board Aircraft

    Committee, 1955), and the other for dynamic elastic

    modulus (Lord and Orkney, 2000). These data are

    0 2 4 6 8 10

    x 105

    0.06

    0.04

    0.02

    0

    0.02

    0.04

    Sensor2response(V)

    20C90C (heating)150C

    Time (s)

    Figure 8. GW signals recorded by Sensor 2 (averaged over30 signals for each temperature) at various temperatures whileheating (offset by a small DC voltage for clarity).

    0 1 2 3 4 50

    50

    100

    150

    Tempera

    ture(C)

    Time (h)

    Figure 7. Typical timetemperature curve for experiments done inthe computer-controlled autoclave.

    1388 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    8/17

    shown in Figure 9. The former was obtained from

    standard stressstrain tests conducted under varying

    temperature for aluminum alloy 7075, while the latter

    was found from measuring changes in natural

    frequency of aluminum beam (alloy 5052) flexural

    vibrations with temperature. No data were found foraluminum 5005, the material used in the tests here.

    However, it is similar in composition to the two

    alloys for which data were found. The variation in

    elastic modulus of PZT-5A is relatively small

    (Williams et al., 2004). No data were obtained for

    dynamic elastic modulus variation of PZT-5A.

    2. Piezoelectric properties of PZT-5A: It is well-known

    that the piezoelectric constants (d31 and g31) vary

    significantly with temperature (Berlincourt et al.,

    2007). For GW SHM, the variation in the product

    d31 g31 is of relevance (the d31 constant is associated

    with actuation shear stress induced, while the g31

    constant is associated with the piezo-sensor sensitiv-ity), and this can vary by as much as 7%, as shown

    in Figure 10. In addition, the dielectric constant of

    PZT-5A increases linearly with temperature, which

    causes the load seen by the function generator to

    increase. This however, does not affect sensor

    response by itself.

    3. Thermal expansion: This is a relatively mild effect,

    and causes the plate thickness, piezo dimensions, and

    distances traveled by the GWs in the plate to

    increase, while material density decreases. Since the

    thermal expansion coefficients of aluminum and

    PZT-5A are known (average values over 20 to

    1508C are 25.5mm/m-8C for aluminum obtainedfrom Matweb (2007); 1 2.5mm/m-8C for PZT-5A,

    see Williams et al., 2004), these effects can be

    accounted for. The effect of changing (static) elastic

    modulus, plate thickness, and density were used to

    compute Lamb-wave phase velocity dispersion curves

    at different temperatures (Figure 11). The dispersion

    curves at equally spaced temperatures are unequally

    spaced, particularly at the higher temperatures. This

    is because of the nonlinear change in Youngs

    modulus and density with temperature.

    4. Damping and pyroelectric effects (not considered)

    Another parameter that changes with increasing

    temperature is damping in the structural substrate

    The best reference found in this regard (Hilton and

    Vail, 1993) estimated an increase by a factor of 4 in

    the loss modulus (representative of damping) a

    100 Hz in aluminum alloy 2024. This still ensure

    that the loss modulus is orders of magnitude lower

    than the elastic modulus and can be neglected, as was

    verified at room temperature in a previous work

    (Raghavan and Cesnik, 2005b). Therefore, damping

    was ignored at higher temperatures too. Finally, due

    to the pyroelectric effect, temperature changes causea static voltage to appear across a piezos electrodes

    Since the experimental signals were acquired with a

    2 Hz high-pass filter built into the oscilloscope, thi

    effect was not considered either.

    Theoretical Modeling and Signal Analysis

    Effects (a)(c) were incorporated into the theoretica

    models developed by the authors (Raghavan and

    Cesnik, 2005b). These models can capture the

    1540

    1560

    1580

    1600

    1620

    1640

    1660

    1680

    0 20 40 60 80 100 120 140 160

    PZT-

    5Ad31xg31

    (C-V/m-sqN)

    Temperature (C)

    Figure 10. Variation of d31g31 of PZT-5A (Berlincourt et al., 2007)

    0

    1000

    2000

    3000

    4000

    5000

    6000

    0 100 200 300 400 500

    Phasevelocity(m/s

    )

    S0 at 20C

    A0 at 150C

    A0 at 80C

    Frequency (kHz)

    A0 at 20C

    S0 at 150C

    S0 at 80C

    Figure 11. Combined effect of changing aluminum elastic modulus(static) and thermal expansion on phase velocity.

    58

    60

    62

    64

    66

    68

    70

    0 20 40 60 80 100 120 140 160

    Youngsmodulus(GPa)

    Al staticAl dynamicPZT-5A

    Temperature (C)

    Figure 9. Variation of Youngs moduli (ANC-5 1955; Lord andOrkney, 2000; Williams et al., 2004).

    Effects of Elevated Temperature on Guided-wave SHM 1389

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    9/17

    three-dimensional GW field excited and sensed by

    arbitrarily shaped, surface-bonded finite-dimensional

    piezos in isotropic plates. The models assume uncoupled

    dynamics and the piezo-actuator is modeled as causing

    surface shear traction along its edge in the direction

    normal to the edge on the surface of the underlying

    substrate. More recently, extensions of these modelsto other structural configurations (isotropic beams,

    hollow cylinders, and composite plates) have also been

    proposed (Raghavan and Cesnik, 2007a,b). These

    models have been validated extensively by numerical

    and experimental results. The inputs for these models

    are the material properties, dimensions of the structure,

    the piezo dimensions, the piezo-actuators shear

    traction magnitude, and piezo-sensors piezoelectric

    constant and Youngs modulus, all of which are

    thermally sensitive parameters as explained above.

    Reduced transducer dimensions were used to generate

    the theoretical plots. A 20% reduction in transducer size

    over nominal values was assumed based on an earlierwork to correlate frequency response plots from

    experiments with theoretical predictions for piezos of

    the same thickness on similar aluminum plates

    (Raghavan and Cesnik, 2005b). This is to account for

    shear lag, which causes strain transfer to happen over a

    small area close to the actuator edge, and not exactly

    the actuator edge, as assumed in the pin-force model.

    In addition, variations in the traction magnitude exerted

    by the piezo due to changing elastic moduli and

    dimensions were accounted for using the equation for

    static actuation by piezos in Chaudhry and Rogers

    (1994). These models were used to generate theoretical

    predictions for the time-domain sensor response signalsat various temperatures between 20 and 1508C for the

    configuration used in the experiments, incorporating

    data for the thermal variation of the model inputs. With

    that, it was examined whether the experimentally

    observed increase in time-of-flight of the GW S0 mode

    wave-packet received by Sensor 2 and the change in

    sensor response peak-to-peak magnitude could be

    captured. Spectrograms of these signals were generated

    (for both the theoretical and experimental ones) and

    the time-of-flight of the first transmitted S0 pulse was

    computed from them. This also allowed examining to

    see whether there was a change in the center frequency

    of this pulse with temperature. The time-of-flight hereis defined as the time corresponding to the peak in the

    first wave pulse (from the spectrogram) minus half

    the excitation pulse span (corresponding to the peak of

    the excitation), as shown in Figure 12. It was calculated

    with a resolution of the experimental sampling rate, i.e.,

    0.1ms. The comparison between theory and experiment

    is shown in Figure 13 for time-of-flight and in Figure 14

    for sensor response magnitude (peak-to-peak). To get

    an estimate of the error in time-of-flight, the raw

    un-averaged 30 signals were collected for two points

    per thermal cycle (1008C while heating and 708C while

    cooling). The data from Sensor 3 is very similar.

    Discussion of Results and Implications for GW SHM

    The theoretical estimates for time-of-flight are in

    agreement with the experimental data (within error

    (a)

    (b)

    (c)

    10

    0

    10

    0.05

    200

    150

    100

    0

    0.05

    Actuation

    signal(V)

    Sensor

    signal(V)

    Frequency

    (kHz)

    Time (s)

    10

    2

    x104

    x104

    0.2 0.4

    Time of flight

    0.6 0.8

    10.2 0.4 0.6 0.8

    x10410.2 0.4 0.6 0.8

    Figure 12. Extraction of time-of-flight at 908C while heating: (a)actuation signal; (b) Sensor 2 response signal; and (c) spectrogramof signal in (b).

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0 20 40 60 80 100 120 140 160

    Cycle 1 heating

    Cycle 1 cooling

    Cycle 2 heating

    Cycle 2 cooling

    Theoretical (static modulus variation)

    Temperature (C)

    Sensor2responsepeak-to-peakmagnitude

    Figure 14. Variation in response magnitude (peak-to-peak) of firsttransmitted S0 mode received by Sensor 2.

    19.5

    20

    20.5

    21

    21.5

    22

    0 20 40 60 80 100 120 140 160

    Time-of-flight(s)

    Exper imental (cycle 1 heat ing) Exper imental (cycle 1 cool ing)

    Exper imental (cycle 2 heat ing) Exper imental (cycle 2 cool ing)

    Theoretical (static elastic modulus) Theoretical (dynamic elastic modulus)

    Temperature (C)

    Figure 13. Variation in time-of-flight of first transmitted S0 modereceived by Sensor 2.

    1390 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    10/17

    margins) and the agreement seems better for the

    theoretical data set generated assuming static elastic

    modulus variation. The center frequency was calculated

    with a resolution of 3 kHz, which is limited by the time

    span of the signal recorded. With this resolution, the

    center frequency of the sensed S0 mode pulse remained

    constant at 133 kHz for both the theoretical andexperimental data sets at different temperatures. For

    the sensor response magnitude prediction, there is

    clearly a significant gap between the theoretical estimate

    and experimental data. The experimentally observed

    increase of up to 33% while heating to 908C in the

    response magnitude of Sensor 2 is not captured by the

    theoretical model, but the decrease of 18% at 1508C is

    predicted within the error margin. One possible expla-

    nation for this is changing bond layer properties with

    temperature (the model used here assumed perfect

    bonding). Data were unavailable for variation of bond

    layer elastic modulus with temperature and therefore it

    could not be quantified. However, the static modulusvariation and thermal expansion data can be used to

    give a good approximation to account for the slowing

    down of wavespeeds with increasing temperature, and

    is used in the subsequent section to generate damage

    location estimates. An empirical compensation

    approach is used for varying response magnitude.

    From these experiments, it is clear that temperature

    can cause significant changes in the magnitude and

    time-of-flight for the first transmitted pulse received by

    the sensor. Even with the instantaneous temperature

    known, this causes greater error margins in the

    magnitude and time-of-flight measurement under

    slightly varying temperature. In most pitch-catchapproaches used, changes in these very features are

    used to conclude whether damage is present in the

    actuatorsensor path or not. It should be clarified that

    in this context, pitch-catch approaches refer to those

    that use a dense network of transducers and rely on

    changes in the GW signal transmitted over the direct

    actuatorsensor paths. Therefore, the pitch-catch

    method is inherently more sensitive to false positives in

    damage detection under varying temperature. On the

    other hand, pulse-echo approaches typically rely on the

    absence or presence of reflected or scattered echoes

    between the actuation/first transmitted pulse and the

    boundary reflection (although some researchers do useechoes arriving after the first boundary reflection).

    If damage is present, regardless of temperature increase,

    there will always be some GW reflection/scattering and

    a sensor adjacent to the actuator or elsewhere should be

    able to pick this up (assuming the damage is sensitive

    to the mode and frequency of the incident GW).

    Therefore, in principle, the only modification to make

    for GW pulse-echo based damage characterization

    under varying temperature would be to account for

    varying GW speeds and scale the reflection magnitude

    according to the changed sensor sensitivity at tha

    particular temperature. However, complications arise

    for pulse-echo methods in structures with features, such

    as rivets in trying to detect and locate mild damage

    which is roughly at the same distance (within a few cm

    from the actuator as the rivet. In this context, damage is

    called mild if the magnitude of the reflected GW by it iscomparable to that of the reflection from the rivet

    structural discontinuity. This is explored experimentally

    in the next section.

    DAMAGE CHARACTERIZATION UNDER

    VARYING TEMPERATURE

    Baseline Signals and Threshold Values

    As mentioned in the previous section, before any

    damage was introduced two data sets were obtained for

    the baseline, pristine condition of the specimen (whichhad transducers bonded with Epotek 353ND) from two

    identical thermal cycles on different days. This was to

    get a sense of the repeatability of the baseline condition

    for a given temperature between the two cycles. The

    baseline signal read by Sensor 1 (collocated with the

    actuator) is shown in Figure 15. There is some non-zero

    signal between the actuation pulse and the boundary

    reflection due to mild A0 mode excitation and some

    reflection from Sensors 2 and 3. At 208C, the A0 mode

    reflection from Sensors 2 and 3 is discernible, while the

    S0 mode reflection is negligible. However, at highe

    temperatures, the S0 mode reflection from Sensors 2

    and 3 becomes stronger (e.g., see Figure 15)Furthermore, there is some error in the magnitudes o

    these mild reflections received at Sensor 1 when data

    from the two cycles are compared. Figure 16 shows one

    of the worse scenarios in this regard, while Figure 17 is

    an example of better repeatability observed among

    2 4 6 8 10

    x 105

    0.1

    0.05

    0

    0.05

    0.1

    Time (s)

    Sensor1respo

    nse(V)

    20C120C (heating)

    S0mode actuation

    Boundary reflections

    Undesired A0

    mode actuation

    A0

    reflectionfrom s2 and s3

    S0

    reflection

    from s2 and s3

    Figure 15. Signal read by Sensor 1 at 20 and 1108C (Cycle 1) fopristine condition.

    Effects of Elevated Temperature on Guided-wave SHM 1391

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    11/17

    the readings. This error in repeatability arises from the

    variation in time-of-flight and sensor response magni-tude under changing temperature. As mentioned earlier,

    there is a 18C error in the temperature read by the

    thermocouples and the autoclave continues heating/

    cooling at 18C/min during the 30-s period of data

    collection for each temperature reading. With large

    temperature-change rates, this problem is further

    exacerbated. This explains why poor signal repeatability

    was observed in the initial experiments in the oven. The

    variation between the two baseline cycles defines the

    threshold values for the subsequent damage detection

    experiments. In practice, it would be advisable to get

    some more data for the variability in baseline condition

    at various temperatures, particularly in less homoge-neous structural layouts with rivets, stiffeners, etc. In

    addition, while data were collected here at 108C

    intervals, it can also be used for baseline interpolation

    at intermediate temperatures. This can be done by

    simple weighted averaging of the two signals taken at

    the multiples of 108C within which the intermediate

    temperature lies (see Figure 18). Of course, this is

    possible because in this case, the variation between

    signals at 108C intervals is not significant enough

    to cause destructive interference between them

    when averaged. However, if the piezo is used to monitorlarger structures or using higher frequencies, the spacing

    between reflections from structural features may

    increase in the time domain and destructive interference

    might become a possibility with weighted averaging. In

    such cases, readings can be taken at more closely spaced

    temperature intervals to overcome this issue.

    Once the two sets of baseline signals were recorded,

    threshold energy levels were obtained from them.

    The threshold energy in this context is defined as the

    peak signal energy obtained from the spectrogram of the

    difference signal between the two baselines (scaled for

    18V actuation as well as to match the peak-to-peak

    magnitude of the first S0 actuation pulse between thetwo) at the same temperature. The threshold energy is

    chosen to be the peak value from the spectrogram in the

    time window between the end of the S0 mode actuation

    signal seen at Sensor 1 and the beginning of the

    boundary reflections (illustrated in Figure 16 and

    Figure 17). As explained in the authors previous work

    (Raghavan and Cesnik, 2007d), some kind of time

    frequency analysis is essential for GW signals because

    they allow for distinguishing among the different GW

    modes possible and also enable tracking the change in

    the center frequency of the scattered GW pulse (since

    damage may be more sensitive to frequencies other than

    the center frequency of the incident GW pulse). This cansignificantly affect the speed of the scattered pulse,

    which can contribute to location estimate errors.

    Indentation Damage

    Subsequently, damage was introduced artificially in

    the plate by drilling. As alluded to at the end of an

    earlier section, the case of mild damage was first

    explored. To ensure that the reflections were not too

    strong, damage was introduced along the axis of weaker

    0 2 4 6 8 10

    105

    0.15

    0.1

    0.05

    0

    0.05

    0.1

    0.15

    Sensor1r

    esponse(V)

    90C85C (Interpolated)80C

    Time (s)

    Figure 18. Illustration of interpolated signal (at 858C) obtained byweighted averaging of recorded signals at 80 and 908C (slightlyoffset by a constant value for clarity) from pristine specimen whileheating during Cycle 1.

    Sensor1

    response(V)

    Frequency

    (kHz)

    0.05

    160

    140

    12010080

    60

    2 4 6

    Time (s)

    Cycle 1 (pristine)Cycle 2 (pristine)Difference

    8

    8 x 103

    6

    4

    2

    10

    2 4 6 8 10

    x 105

    x 105

    180

    0.05

    0

    Thresholdenergyvalue

    Time-frequencywindow forthresholdenergy

    Figure 17. Sensor 1 response during Cycles 1 and 2 for pristinecondition along with the difference signal and its spectrogram at 608Cwhile cooling (signals offset by a small DC voltage for clarity).

    Sensor1

    response(V)

    Frequency

    (kHz)

    0.05

    0.05

    0

    150

    50

    100

    2 4 6 8 10

    0.15

    Cycle 1 (pristine)Cycle 2 (pristine)Difference

    0.05

    0

    0.1

    x 105

    2 4 6

    Time (s)

    8 10

    x 105

    Time-

    frequencywindow forthresholdenergy

    Thresholdenergyvalue

    Figure 16. Sensor 1 response during Cycles 1 and 2 for pristinecondition along with the difference signal and its spectrogram at1208C while heating (signals offset by a small DC voltage for clarity).

    1392 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    12/17

    actuation of the rectangular actuators. Theoretical

    models developed earlier by the authors (Raghavan

    and Cesnik, 2005b) predicted a 30% weaker GW strain

    field along this direction. Experiments were first donefor a triangular cross-sectional indentation of maximum

    diameter 5 mm and depth 1.7 mm (Figure 19(a)). The

    damage center is 8 cm away from the plate center and its

    location is shown in Figure 4. The signal obtained from

    Sensor 1 at room temperature after introducing this

    damage is shown in Figure 20, along with the pristine

    condition signal and the difference between the two.

    As a well-established practice in the GW SHM commu-

    nity, this difference signal is used for damage detection

    and characterization. There are clear reflections in the

    difference signal whose peak energies from the spectro-

    gram are well above the energy threshold defined for

    208C. The two distinct reflections in the signal are theS0- and A0-mode reflections from the indentation.

    As explained in the authors earlier work (Raghavan

    and Cesnik, 2007d), even though the S0 mode is

    predominantly excited, mode conversion is possible at

    a defect. In this case, because of the controlled

    introduction of damage in the specimen, the A0 and S0mode reflections were expected and their time-of-flight

    values could be estimated. However, in field applica-

    tions, one has to blindly infer all information from the

    GW signals. Therefore, for practical situations, one has

    to be able to identify the GW mode of each reflection as

    well. This can be done from the reflections character-

    istics in the timefrequency plane. For the A0 mode,since higher frequencies travel much faster than the

    lower ones, the peaks at higher frequencies within the A0mode reflection arrive faster than those for the lower

    ones in the timefrequency plane. On the other hand,

    for the S0 mode the converse is true, but the variation in

    S0 mode wavespeeds is milder over the frequency range

    and structure used in these experiments. Once the mode

    is identified, the group speed(s) corresponding to the

    frequency at which the peak within the reflection occurs

    is used with the time-of-flight for estimating the radial

    location of the damage site (see Raghavan and Cesnik

    2007d). As mentioned in an earlier section, the group

    speeds for a particular temperature are computed by

    accounting for the Youngs modulus at that temperatureand thermal expansion in the dispersion curves. In this

    case, the GW packet originates from one edge of the

    actuator and its echo travels back all the way to the

    other edge and has to further traverse a distance equal to

    half the sensor dimension before the peak of the

    reflected GW is seen in the sensor response. Therefore

    minor correction terms to account for the transi

    times of the GW packet through the transducers are

    subtracted from these estimates. At 208C, the radia

    location estimates of the indentation are 7.2 and 8.2 cm

    (from the plate center) based on the S0 and A0 mode

    reflections, respectively. The analysis done in thi

    study was done by manually tracking the reflections inthe spectrograms. In principle, the chirplet matching

    pursuit algorithm used earlier by the authors (Raghavan

    and Cesnik, 2007d) should allow automated tracking

    and better resolution for real-time processing in

    end applications. However, its implementation in

    LastWave 2.0 (Bacry, 2007) makes approximations to

    reduce computational complexity which do no

    capture the slowing down of wavespeeds at elevated

    temperatures seen in Figure 13.

    The indented specimen was thermally cycled in the

    autoclave to check whether the difference signa

    remained above the pre-defined threshold level at each

    temperature point. Some of the signals read by Sensor 1during this experiment are shown in Figure 20 and

    Figure 21. The results are also summarized in Table 1

    The A0 and S0 mode reflections from the indentation

    had peak energy (from the spectrogram, in the excited

    frequency band as shown in Figure 20 for the signal at

    208C) well above the threshold up to 808C while heating

    Some of the S0 mode reflections (from 50 to 808C) mixed

    with the excitation difference signal, due to which the

    S0 mode reflection underestimated the damage location

    (shown in Figure 21). Beyond 808C, the A0 mode

    (a) (b)

    Figure 19. Photographs of damage introduced: (a) indentation and(b) through-hole.

    Sensor1

    response(V)

    Frequency

    (kHz)

    0.05

    0

    2 4 6 8 10

    2 4 6

    Time (s)

    PristineIndentation

    Difference

    8 10

    0.2

    0.1

    0

    x 105

    x 105

    0.05

    180

    160140

    120

    100

    80

    60

    S0

    modereflection

    A0

    modereflection

    Figure 20. Sensor 1 response at 208C for pristine and indented specimens, along with the difference signal (offset by smaDC values for clarity) and its spectrogram.

    Effects of Elevated Temperature on Guided-wave SHM 1393

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    13/17

    Table

    1.

    Summaryo

    fresu

    lts

    show

    ing

    trends

    in

    thermalexperimentfordamage

    cha

    racterization

    with

    indented

    specimen

    (Key:

    Temp.

    Temperature;

    (h)

    heating

    phase;

    (c)

    cooling

    phase;

    S0

    ToF

    time-o

    f-flightfor

    S0

    mode

    reflection

    from

    indentation;

    S0

    loc.

    est.

    radiallocatione

    stimate(relativetoplatecenter)ofdamagebasedonS

    0

    modereflectio

    n).

    Temp.

    (8C)

    S0

    ToF(ks)

    S0

    energy(V2)

    Threshold

    energy(V2)

    S0

    above

    threshold?

    S0

    loc.est.

    (cm)

    A0

    ToF(ks)

    A0

    energy(V2)

    A0

    above

    threshold?

    A0loc.est.(cm)

    20

    (h)

    28.1

    0

    0.0

    631

    0.0

    142

    Yes

    7.2

    50.6

    5

    0.2

    8825

    Yes

    8.2

    40

    (h)

    28.0

    0

    0.1

    229

    0.0

    025

    Yes

    7.2

    51.2

    0

    0.3

    695

    Yes

    8.6

    70

    (h)

    18.6

    0

    0.0

    739

    0.0

    154

    Yes

    4.7

    54.6

    0

    0.5

    895

    Yes

    9.1

    110

    (h)

    31.2

    0

    0.1

    808

    0.1

    850

    No

    7.9

    56.0

    0

    0.1

    959

    Yes

    9.2

    130

    (h)

    18.6

    0

    0.0

    66

    0.0

    517

    Yes

    4.5

    34.1

    0

    0.0

    886

    Yes

    4.9

    150

    0.0

    275

    25.5

    0

    0.0

    839

    Yes

    4.0

    120

    (c)

    0.0

    375

    28.6

    0

    0.1

    363

    Yes

    4.1

    80

    (c)

    25.5

    0

    0.2

    688

    0.0

    525

    Yes

    6.5

    56.8

    0

    0.1

    502

    Yes

    9.5

    50

    (c)

    27.5

    0

    0.0

    681

    0.0

    033

    Yes

    7.1

    55.8

    2

    0.2

    240

    Yes

    9.4

    20

    (c)

    30.3

    5

    0.0

    483

    0.0

    023

    Yes

    7.8

    53.9

    5

    0.4

    228

    Yes

    8.8

    1394 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    14/17

    reflection was still above the threshold, but had peakenergy that was comparable to the threshold. The

    weaker S0 mode reflection had peak energy lower than

    the threshold at some points after 908C while heating.

    In addition, as illustrated in Figure 21, at some points

    over 1008C, there is a reflection that arrives approxi-

    mately where the S0 mode was seen at lower tempera-

    tures, but can be wrongly identified as A0 mode by its

    timefrequency characteristics. This results in much

    larger errors for the radial location estimate.

    Subsequently, while cooling below 808C, again both

    reflections were well above the threshold, and gave

    reasonable location estimates.

    Through-hole Damage

    Following this, a through hole of diameter 7.5 mm

    was drilled at the same location (Figure 19(b)). The

    response of Sensor 1 recorded after drilling through at

    208C is shown in Figure 22. As seen in the difference

    signal, the S0 mode reflection is now much stronger

    while the A0 mode reflection appears much weaker

    (but still well above the threshold at 208C). The radial

    location estimates are 8.6 and 8.3 cm based on the S0 and

    A0 mode reflections, respectively at 208C. The results

    from thermally cycling the specimen with the through-

    hole are shown in Figures 22 and 23 and tabulated inTable 2. In this case, the S0 mode reflection is reasonably

    above the threshold at all temperatures. There is an

    error of 1.9 cm in the estimate (based on the S0 mode) at

    1508C, where the signal is the weakest. At a couple of

    points, the radial location of the damage is over-

    estimated by more than 1 cm (1008C while heating and

    508C while cooling), which is possibly due to the mixing

    of the S0 mode with the difference in the reflection from

    Sensors 2 and 3. At other temperatures, the radial

    location estimates based on the S0 mode are within 1 cm.

    The A0 mode reflection, which was weak at 208C to

    begin with, is discernible up to 708C while heating, bu

    beyond that is indistinguishable from the difference in

    boundary reflection until the specimen cools back to

    room temperature.

    DISCUSSION

    Thus, for mild damage up to 808C, detection was no

    problematic, but there was a slightly increased error in

    location as temperature increased. However, beyond

    that temperature, there is a definite decrease in

    sensitivity, as reflected in the poorer detection

    characterization capability in the indentation experi

    ment. This can be attributed to the higher sensitivity

    of the substrate elastic modulus to temperature

    at higher temperatures (Figure 9), causing greater

    variation in the mild reflections from Sensors 2 and 3

    2 4 6 8 10

    (a)

    Pristine

    Indentation

    Difference

    1 2 3 4 5 6 7 8 9 10 11

    x 105

    x 10

    5

    0.1

    0.05

    0

    0.05

    0.1

    0.1

    0.05

    0

    0.05

    0.1

    Pristine

    Indentation

    Difference

    Sensor1

    response(V)

    S0

    reflection mixed with excitation

    A0

    reflection

    Sensor1

    response(V)

    A0

    reflectionnegligible

    Time (s)

    (b)S0

    reflection appearslike A

    0mode in t-f ane

    Figure 21. Sensor 1 response for pristine and indented specimens,along with the difference signal (offset by small DC values for clarity)at (a) 608C during heating and (b) 1508C.

    0.1

    0.05

    0

    0.05

    2 4 6 8 10

    0

    (b)

    PristineHoleDifference

    PristineHoleDifference

    0.05

    0.05

    0.1

    x 105

    2 4 6 8 10

    (a) x 105

    Figure 23. Sensor 1 response during heating for pristine andthrough-hole specimens, along with the difference signal (offset bysmall DC values for clarity) at: (a) 70 and (b) 1408C.

    Sensor1

    response(V)

    Frequency

    (kHz)

    0.05

    0.05

    0

    S0

    modereflection

    A0

    modereflection

    Pristine

    Hole

    Difference

    Time (s)

    280

    100

    120

    140

    160

    4 6 8 100

    0.05

    0.1

    0.15

    0.2

    2 4 6 8 10

    x 105

    x 105

    Figure 22. Sensor 1 response at 208C for pristine and through-hole specimens, along with the difference signal (offset by smaDC values for clarity) and its spectrogram.

    Effects of Elevated Temperature on Guided-wave SHM 1395

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    15/17

    Table2.

    Summaryofresu

    ltsshow

    ingtrendsinthermalexperiment

    fordamagecharacterizationusings

    pecimenw

    iththrough-h

    ole(Key:

    Te

    mp.

    Temperature;

    (h)

    heatingphase;

    (c)

    coolingphase;

    S0

    ToF

    time-o

    f-flightfo

    rS

    0

    modereflectionfrom

    thru-h

    ole;

    S0

    loc.

    est.

    radiallocationestima

    te(relativetoplate

    center)ofdamagebasedonS

    0

    modereflection;

    Bndry

    A0

    mode

    reflectionpeakw

    ithinboundaryreflection).

    Temp.

    (8C)

    S0

    ToF(ks)

    S

    0

    energy(V2)

    Threshold

    energy(V2)

    S0

    above

    threshold?

    S0

    loc.est.(cm)

    A0

    ToF

    (ks)

    A0

    energy(V2)

    A0

    above

    threshold?

    A0

    loc.est.(cm)

    20

    (h)

    33.4

    5

    0.2

    198

    0.0

    142

    Y

    es

    8.6

    49.7

    5

    0.0

    763

    Yes

    8.3

    40

    (h)

    34.3

    0

    0.2

    972

    0.0

    025

    Y

    es

    8.8

    55.2

    0

    0.1

    119

    Yes

    9.0

    70

    (h)

    32.5

    0

    0.2

    463

    0.0

    154

    Y

    es

    8.3

    56.7

    0

    0.2

    128

    Yes

    9.0

    110

    (h)

    40.6

    0

    0.2

    578

    0.1

    850

    Y

    es

    10.2

    Bnd

    ry

    130

    (h)

    33.4

    0

    0.3

    105

    0.0

    517

    Y

    es

    8.2

    Bnd

    ry

    150

    25.6

    0

    0.1

    073

    0.0

    275

    Y

    es

    6.1

    Bnd

    ry

    120

    (c)

    31.0

    0

    0.1

    997

    0.0

    375

    Y

    es

    7.6

    Bnd

    ry

    80

    (c)

    28.0

    0

    0.2

    985

    0.0

    525

    Y

    es

    7.0

    Bnd

    ry

    50

    (c)

    36.2

    0

    0.3

    393

    0.0

    033

    Y

    es

    9.3

    Bnd

    ry

    20

    (c)

    33.8

    0

    0.2

    018

    0.0

    023

    Y

    es

    8.7

    53.6

    0

    0.0

    623

    Yes

    8.8

    1396 A. RAGHAVAN AND C. E. S. CESNIK

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    16/17

    and consequently, poorer repeatability of the

    baseline signals. In addition, the sensor sensitivity

    (in terms of signal magnitude) drops below the value

    at 208C beyond 1308C (as seen in Figure 14). For the

    through-hole, which can be termed moderate damage,

    detection was clearly possible at all temperatures, but

    at a few points (3 of a total of 29 cases), there wasinaccuracy in the location estimate (by up to 2.2 cm for

    damage located at 8 cm), partly due to interference with

    the difference in reflection from Sensors 2 and 3

    (simulating structural features in field applications).

    One way to reduce the error in location is to use higher

    center frequencies and/or fewer number of cycles (which

    increases frequency bandwidth) in the actuation signal.

    However, that was not feasible in the present setup

    (which did not use amplifiers), since significant distor-

    tion was observed for such actuation signals at higher

    temperatures due to increasing actuator capacitance.

    In addition, in the present experimental setup, the

    scaling of the signals to compensate for changingactuation level increased the signal-to-noise ratio at

    higher temperatures. This indicates the need for reliable

    actuation signal amplification for actuating piezos at

    elevated temperatures.

    SUMMARY AND CONCLUSIONS

    This study addressed the issue of GW SHM using

    piezos under thermal variations as expected in spacecraft

    internal structures. Experiments were done to determine

    a bonding agent (for piezos on aluminum plates) that

    did not degrade under thermal variations from 20 to1508C. Using this bonding agent (Epotek 353ND),

    results from controlled experiments done to examine

    changes in GW propagation and transduction using

    PZT-5A piezos under quasi-statically varying tempera-

    ture (also from 20 to 1508C) were presented. Thermally

    sensitive variables in the experiments were identified and

    quantified to model the experimentally observed

    changes under temperature variation. The Youngs

    modulus of the structure was the most critical of these

    parameters. The increase in time-of-flight of GW pulses

    with increasing temperature was captured by the model

    (within the error margins). However, there was a

    significant gap in the prediction of the large increase insensor response magnitude up to 1008C. The stronger

    vulnerability of pitch-catch approaches (that detect

    damage in the direct actuatorsensor path based on

    changes in the first pulse transmitted over this path) to

    false positives under changing temperature was then

    explained. Finally, detection and location of damage

    (by drilling) using the pulse-echo GW testing approach

    in the presence of mild structural GW scatterers

    (to simulate rivets) was explored in the same tempera-

    ture range. Damage characterization of a half-plate

    thickness indentation at 8 cm from the actuators was no

    significantly affected up to 808C, but beyond tha

    temperature, detection/characterization was difficult

    The problems beyond 808C can be traced to increased

    sensitivity of substrate elastic modulus to temperature

    and weaker sensor sensitivity beyond 1308C. For a

    through-hole, damage detection and characterizationwas possible at all temperatures and, except at a few

    temperatures (3 out of 29), damage was located within

    1 cm for a nominal location of 8 cm and, hole diamete

    0.75cm. Suggested approaches for improving the

    sensitivity at higher temperatures include testing a

    higher frequencies and/or with shorter time-span

    excitation pulses, with reliable, low-distortion actuation

    amplifiers.

    ACKNOWLEDGMENTS

    The authors gratefully acknowledge Dr. W.H. Prosser

    (NASA Langley) for suggesting the investigation o

    b o nd in g agen t d egr ad ati on . T he ass i st ance o

    Mr. Danny Lau (University of Michigan) in setting up

    one of the experiments is also appreciated. This work

    was supported by the Space Vehicle Technology

    Institute under grant NCC3-989 jointly funded by

    NASA and DoD within the NASA Constellation

    University Institutes Project, with Ms. Claudia Meye

    as the project manager.

    REFERENCES

    Bacry, E. 2007. LastWave 2.0 software, http://www.cmap.polytechniquefr/$ bacry/LastWave

    Berlincourt, D., Krueger, H.H.A. and Near, C. 2007. Properties oMorgan Electroceramics, Morgan Electroceramics TechnicaPublication TP-226, http://www.morganelectroceramics.com

    Blaise, E. and Chang, F.-K. 2001. Built-in Diagnostics foDebonding in Sandwich Structures under ExtremeTemperatures, Proceedings of the 3rd International Workshopon Structural Health Monitoring, pp. 154163, StanfordCalifornia, USA.

    Carlos, G.-C., Alan, C. and Gordon, A. 2005. Health Managemenand Automation for Future Space Systems, Proceedings of the

    AIAA Space 2005 Conference and Exposition, pp. 15411557Long Beach, CA.

    Chambers, J., Wardle, B. and Kessler, S. 2006. DurabilityAssessment of Lamb-wave Based Structural Health MonitoringNodes, Proc. 47th AIAA/ASME/ASCE/AHS/ASC StructuresStructural Dynamics, and Materials Conference, Paper # AIAA2006-2263, Newport, Rhode Island.

    Chaudhry, Z. and Rogers, C.A. 1994. The Pin-force ModeRevisited, Journal of Intelligent Material Systems andStructures, 5(3):347354.

    Derriso, M.M., Olson, S., Braisted, W., DeSimio, M., Rosenstengel, Jand Brown, K. 2004. Detection of Fastener Failure in aThermal Protection System, Proceedings of the SPIE, 5390585596.

    Effects of Elevated Temperature on Guided-wave SHM 1397

    by guest on October 5, 2010jim.sagepub.comDownloaded from

    http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/http://jim.sagepub.com/
  • 8/6/2019 Effects of Elevated Temperature on Guided-Wave

    17/17

    Ecke, W., Latka, I., Willsch, R., Reutlinger, A. and Graue, R. 2001.Fiber Optic Sensor Network for Spacecraft HealthMonitoring, Measurement Science and Technology, 12:974980.

    Epoxies Etc. 2006. 10-3004 Adhesive Technical Datasheet, http://www.epoxies.com, Cranston, RI.

    Epoxy Technology 2006. Epoxies 301 and 353ND TechnicalDatasheets, http://www.epotek.com, Billerica, MA.

    Friebele, E.J., Askins, C.G., Bosse, A.B., Kersey, A.D., Patrick, H.J.,

    Pogue, W.R., Putnam, M.A., Simon, W.R., Tasker, F.A.,Vincent, W.S. and Vohra, S.T. 1999. Optical Fiber Sensorsfor Spacecraft Applications, Smart Materials and Structures,8:813838.

    Hilton, H.H. and Vail, C.F. 1993. Bending-torsion Flutter of LinearViscoelastic Wings Including Structural Damping, Proc.AIAA/ASME Structures, Structural Dynamics and MaterialsConference, Paper # AIAA-93-1475-CP, La Jolla, CA.

    Huang, J., Rose, J., Gordon, J. and Boucher, R. 2005. StructuralSensor Testing for Space Applications, Proceedings of the SPIE,5762:8899.

    Konstantinidis, G., Drinkwater, B.W. and Wilcox, P.D. 2006. TheTemperature Stability of Guided Wave Structural HealthMonitoring Systems, Smart Materials and Structures,15:967976.

    Larson, W.J. and Wertz, J.R. 1995. Space System Design and Analysis,

    2nd edn, Microcosm, CA/Kluwer Academic, The Netherlands.Lee, B.C., Manson, B. and Staszewski, W.J. 2003. Environmental

    Effects on Lamb Wave Responses from Piezoceramic Sensors,Materials Science Forum, 440441:195202.

    Lin, M., Kumar, A., Qing, X., Beard, B.J., Russell, S.S., Walker, J.L.and Delay, T.K. 2003. Monitoring the Integrity of FilamentWound Structures by Built-in Sensor Networks, Proceedings ofthe SPIE, 5054:222229.

    Lord, J.D. and Orkney, L.P. 2000. Elevated Temperature ModulusMeasurements using the Impulse Excitation Technique,National Physical Laboratory Measurement Note,Teddington, UK.

    Lu, Y. and Michaels, J.E. 2005. A Methodology for StructuralHealth Monitoring with Diffuse Ultrasonic Waves in thePresence of Temperature Variations, Ultrasonics, 43:713731.

    Matweb 2007. Material property database, http://www.matweb.com

    Myers, D.E., Martin, C.J. and Blosser, M.L. 2000. ParametricWeight Comparison of Advanced Metallic, Ceramic Tile, andCeramic Blanket Thermal Protection Systems, NASA-TM-2000-210289, NASA Center for AeroSpace Information.

    Raghavan, A. and Cesnik, C.E.S. 2005a. Lamb Wave Methods inStructural Health Monitoring, In: Inman, D., Farrar, C.R.,Lopes, Jr, V. and Steffen, Jr, V. (eds), Damage Prognosis, JohnWiley & Sons, UK.

    Raghavan, A. and Cesnik, C.E.S. 2005b. Finite DimensionalPiezoelectric Transducer Modeling for Guided Wave BasedStructural Health Monitoring, Smart Materials and Structures,

    14:14481461.

    Raghavan, A., and Cesnik, C.E.S. 2007a. 3-D Elasticity-basedModeling of Anisotropic Piezocomposite Transducers forGuided-wave Structural Health Monitoring, Journal of

    Vibration and Acoustics (Transactions of the ASME), SpecialIssue on Damage Detection and Structural Health Monitoring129(6):739751.

    Raghavan, A. and Cesnik, C.E.S. 2007b. Modeling of Guided-waveExcitation by Finite-Dimensional Piezoelectric Transducers inComposite Plates, Presented at the 15th AIAA/ASME/ASCAdaptive Structures Conference, Paper 2007-1725, Honolulu,Hawaii, Apr. 2326.

    Raghavan, A. and Cesnik, C.E.S. 2007c. Review of Guided-waveStructural Health Monitoring, The Shock and Vibration Digest,39:91114.

    Raghavan, A. and Cesnik, C.E.S. 2007d. Guided-wave SignalProcessing using Chirplet Matching Pursuits and Mode

    Correlation for Structural Health Monitoring, SmartMaterials and Structures, 16:355366.

    Schulz, M.J., Sundaresan, M.J., Mcmichael, J., Clayton, D., Sadler, R.and Nagel, B. 2003. Piezoelectric Materials at ElevatedTemperature, Journal of Intelligent Material Systems andStructures, 14(11):693705.

    Stanley, D., Cook, S. and Connolly, J. (study managers) 2005.Exploration Systems Architecture Study Final Report, NASA-TM-2005-214062, NASA Space Transportation InformationNetwork.

    U.S. Munitions Board Aircraft Committee 1955. Strength ofMetal Aircraft Elements, ANC-5 Bulletin, Republican Press,Hamilton, Ohio.

    Williams, B.R., Inman, D.J. and Wilkie, W.K. 2004.Temperature-dependent Thermoelastic Properties for MacroFiber Composite Actuators, Journal of Thermal Stresses,

    27:903915.Yang, J., Chang, F.-K. and Derriso, M.M. 2003. Design of a Built-in

    Health Monitoring System for Thermal Protection Panels,Proceedings of the SPIE, 5046:5970.

    1398 A. RAGHAVAN AND C. E. S. CESNIK