Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Engineering Structures 31 (2009) 2224–2235

Contents lists available at ScienceDirect

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

Real-time wireless vibration monitoring for operational modal analysis of an


integral abutment highway bridge
Matthew J. Whelan a , Michael V. Gangone a , Kerop D. Janoyan b,∗ , Ratneshwar Jha c
a
Clarkson University, Department of Civil and Environmental Engineering, Potsdam, NY 13699-5712, United States
b
Clarkson University, Department of Civil and Environmental Engineering, Potsdam, NY 13699-5710, United States
c
Clarkson University, Department of Mechanical and Aeronautical Engineering, Potsdam, NY 13699-5725, United States

article info abstract


Article history: Remote structural health monitoring systems employing a sensor-based quantitative assessment of in-
Received 8 May 2008 service demands and structural condition are perceived as the future in long-term bridge management
Received in revised form programs. However, the data analysis techniques and, in particular, the technology conceived years ago
17 March 2009
that are necessary for accurately and efficiently extracting condition assessment measures from highway
Accepted 19 March 2009
Available online 7 May 2009
infrastructure have just recently begun maturation. In this study, a large-scale wireless sensor network
is deployed for ambient vibration testing of a single-span integral abutment bridge to derive in-service
Keywords:
modal parameters. Dynamic behavior of the structure from ambient and traffic loads was measured
Bridge dynamics with accelerometers for experimental determination of the natural frequencies, damping ratios, and
Modal analysis mode shapes of the bridge. Real-time data collection from a 40-channel single network operating with a
Bridge inspection sampling rate of 128 Hz per sensor was achieved with essentially lossless data transmission. Successful
Wireless sensor network acquisition of high-rate, lossless data on the highway bridge validates the proprietary wireless network
Structural health monitoring protocol within an actual service environment. Operational modal analysis is performed to demonstrate
Stochastic subspace identification (SSI) the capabilities of the acquisition hardware with additional correlation of the derived modal parameters
to a Finite Element Analysis of a model developed using as-built drawings to check plausibility of the
mode shapes. Results from this testing demonstrate that wireless sensor technology has matured to the
degree that modal analysis of large civil structures with a distributed network is a currently feasible and
a comparable alternative to cable-based measurement approaches.
© 2009 Elsevier Ltd. All rights reserved.

1. Introduction of structural retrofitting. Additionally, inherent monitoring of


environmental loads and influences, operational loads, and traffic
Highway administrators have been faced with the burden of patterns and densities can be used to collect a database of field
managing a rapidly aging network of highway bridges in which measurements for providing feedback on bridge design practice.
a significant portion have met or exceeded their design lifetime A study of over 500 bridge failures conducted by Wardhana
and service limits. As demonstrated in the aftermath of recent and Hadipriono [1] assessing events from 1989 to 2000 concluded
bridge collapses over the past several decades, current schedule- that the majority of collapse instances occur due to a triggering
based visual inspections fall short of ensuring a safe operational event. In particular, short-term hydraulic events, long-term scour,
model for highway bridge management with bridge closures collision, and overload were sighted for 73% of the documented
preceding imminent failure. Following development of advanced collapse, while deterioration of structural members, design flaws,
diagnostic and prognostic approaches, in-service monitoring of and construction-related issues resulted in nearly 12% of the
highway bridges with sensor networks may serve to evaluate failures. Collisions, scour, and structural deterioration significant
the operational health of a particular structure and estimate enough to produce bridge collapse should produce detectable
the remaining service life. Wireless sensor networks furthermore changes in the dynamic response of the structure. Feedback from a
enable the rapid instrumentation of bridges for assessing the sensor-based monitoring system would preemptively signal such
impact of construction-related activities and evaluating the effect deterioration to permit a schedule of repair or closure prior to
unsafe operation. The aforementioned study also noted that bridge
failures due to overloading were the most devastating in terms
∗ Corresponding address: Clarkson University, Department of Civil and Environ- of human casualties. Since the load carrying capability is not a
mental Engineering, 8 Clarkson AveBox 5710, 13699-5710 Potsdam, NY, United
static quantity over its service life, routinely assessing and posting
States. Tel.: +1 (315) 268 6506; fax: +1 (315) 268 7985. the structural capacity of bridge structures is further essential for
E-mail address: kerop@clarkson.edu (K.D. Janoyan). maintaining public safety.
0141-0296/$ – see front matter © 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2009.03.022
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2225

Due to the reasonable limitation on the density of deployed diaphragms and two equally spaced C15 × 33.9 intermediate
sensors across large civil structures, nondestructive testing has diaphragms between all girders (Fig. 1). The abutments are a
generally focused on characterization of changes in global dynamic U-type integral design supported by nine HP10 × 42 piles with
properties as a consequence of local damage. A limited library of strong-axis orientation at the south abutment and weak-axis
vibration-based instrumentation studies exist that have measured orientation at the north abutment. The bridge has an inventory
modal parameters of a full-scale bridge prior to and after load rating of 40.8 metric tons for the HS25.4 truck and an
progressive induced damage scenarios [2–5]. However, a definitive operational load rating of 68.9 metric tons for the HS42.4, as
method of deriving damage identification and localization has determined through the load factor (LF) method. The bridge was
yet to be formulated. Consequent to the prohibitively high cost inspected by a Professional Engineer through the NYSDOT under
and oversight associated with permanent, continuous monitoring the National Bridge Inspection Standards (NBIS) in the same month
of a county or statewide distributed system of bridges, short- as the in-service monitoring. Given that the bridge was only two
term monitoring, concurrent to existing visual inspections, has years old at the time of inspection, the deck, superstructure, and
also been another, more utilized, approach. As an overwhelming substructure were all assigned condition ratings of 8 on the NBIS
number of bridges nationwide have approached or surpassed their 0–9 rating scale.
service lifespan, it has become critical to allocate resources to
the structures most in need of repair or replacement. Diagnostic 3. Test setup, method, and instrumentation
testing through short-term strain measurements from known
A total of 20 data points were monitored with dual-axis
truck load patterns has been utilized to assess the condition of
accelerometers, as shown in Fig. 2. The instrumentation layout
these bridges [6]. This approach generally merges the results of the
was chosen to balance the requirements of the modal testing
experimental strain analysis with codified analytical load rating
outlined and a concurrent quasi-static investigation using strain
measures to determine revised operating loads and ratings for the
transducers connected to the same wireless hardware. The general
in-service bridge. The hardware deployed in this study is unique
test setup consists of twenty dual-axis accelerometers deployed
in that it provides a wireless platform for both vibration response alongside eleven reusable strain transducers that were interfaced
measurement as well as static load ratings through strain readings. with twenty wireless sensor nodes. All wireless sensor nodes
While traditional cable-based sensors can be utilized to assist communicated in a single-network star topology with a central
the field engineer with schedule-based inspections, the excessive coordinator node connected to a local microcomputer notebook.
instrumentation costs and time of installation often limits their Acceleration responses of the bridge from ambient excitation
use to special cases. Recently, there has been much interest in were measured using low-cost Micro-Electro-Mechanical Systems
the use of wireless transceivers for communication of sensor (MEMS) accelerometers mounted directly to the web of each
data to alleviate the burdens associated with widely-distributed girder using a fast curing epoxy. All measurement data was
cable-based sensors [7]. However, while the number of unique wirelessly streamed to the microcomputer in real-time in an
wireless sensor platforms has continued to rapidly expand, there acceleration time history format. The acquisition of complete
has been limited success in replicating previous cable-based test time histories, rather than preprocessed spectra or extracted
programs in regard to the number of deployed sensors and data parameters, permits the application of multiple post-processing
acquisition rates. A review of recent wireless sensor deployments analysis methods including those necessitating multi-node time-
for structural health monitoring of bridges [8–10] reveals that the series data or high computational overhead, such as stochastic
networks have generally relied on either local data logging and subspace identification.
post-sampling transmission of sensor data or on low sampling
rates and/or limited numbers of sensors in order to address 3.1. Test methodology
transceiver bandwidth limitations. Such concessions severely limit
the versatility and capability of a structural health monitoring Individual tests consisted of time history responses concur-
system in terms of sampling duration, data acquisition rates, and rently streamed in real-time from all forty sensor channels dis-
spatial resolution as well as quality of the derived mode shapes. tributed across twenty wireless sensor units. Durations of three
The wireless sensor network deployed in this study has minutes and six seconds (186 s) were specified with an effective
achieved more than adequate sampling rates necessary for sampling rate of 128 Hz in an effort to replicate similar recent
modal analysis of highway bridges, including short-span and stiff experimental programs utilizing wired instrumentation as doc-
structures, while maintaining reliable communication within a umented in Wenzel and Pichler [11]. Excitation of the structure
large, dense array of sensors. Consequent to years of development was provided only by ambient environmental loads and vehicular
and limited field success with high throughput networks, wireless traffic. The bridge selected exhibits very low-level vibration from
sensor technology has often become viewed as a conceptually ambient loading due to its relatively short-span length and the
ideal solution to the problem of in-service structural condition integral abutment design, thereby providing a challenging plat-
assessment, although incapable of providing the instrumentation form for structural response measurement and modal analysis.
The accelerometers were distributed along all of the girders
framework needed. The results of this paper aim to demonstrate
in a pattern (Fig. 2) established to derive the operational
that low-cost wireless technology has emerged to the point that
mode shapes. Finite element analysis was performed in advance
it is now a feasible and comparable alternative to cable-based
to ensure that data points did not correspond with nodes
structural instrumentation systems.
of zero displacement for the modes of interest. One central
girder was instrumented with a denser array as a result of
2. Test structure concurrent strain monitoring; this configuration resulted in
increased spatial resolution along this girder. The vertical and
The in-service bridge investigated consists of a 24.1 cm (9.5 in) longitudinal vibration responses were measured as a result of
thick reinforced concrete slab supported by four integral abutment sensor orientation. Longitudinal responses were inadvertently
girders with a single-span of 17.07 m (56 ft). Carrying Wright Road measured due to the orientation of the sensor placement.
over Trout Brook in Potsdam, N.Y., the bridge was constructed Transmission of the longitudinal acceleration served solely to
in 2004 and is under the jurisdiction of the St. Lawrence County demonstrate the network throughput capability. Consequently,
Department of Highways (NBI# 000000003231620). Four W36 × only the vertical vibration response was incorporated in the
135 steel beams are spaced at 2.74 m (9 ft) with MC8 × 20 end operational modal analysis.
2226 M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235

Fig. 1. Concrete slab on steel girder integral abutment bridge.

Fig. 2. Vibration measurement locations.

Capacitance-based dual-axis MEMS accelerometers (STMicro- accelerometer channels depending on sensor location to maximize
electronics LIS2L02AL) were selected for low-power consumption the range and resolution of the conversion. This signal condition-
and ultra-low noise floor characteristics. These integrated circuit ing resulted in approximate sensitivities of 38 Vg−1 and 76 Vg−1 ,
sensors were mounted on printed circuit boards and encased in a respectively, with analog-to-digital conversion resolution of 16 µg
small external housing with potting epoxy to enable direct place- and 8 µg, accordingly.
ment of the sensor on the structure for superior vibration transfer. Conversion of the raw signals is provided at each node by a
−1
The accelerometers feature +/ − 2 g full-scale range,
√ 600 mV g 12-bit ADC (Analog-to-Digital Converter) that was programmed to
sensitivity at 3V supply, and an ultra-low 30 µg/ Hz noise den- over-sample each measurement channel at 512 Hz. The data was
sity. Signal amplification of 64V/V or 128V/V was applied to the then processed using a 56th-order digital low-pass filter to remove
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2227

potential alias frequencies and then down-sampled to an effective Data packets from each node are scheduled for transmission
rate of 128 Hz. The oversampling approach implemented increases based on the local address of the node to prevent packet
the effective resolution of the ADC, rejects alias frequencies from collision. The built-in clear channel assessment (CCA) feature
the transition band of the analog low-pass filter, and produces a of the transceiver is also utilized to further prevent packet
near brick-wall filter response with less than 0.01 dB of attenuation transmission when the channel is already active. Data packets
in the 0–50 Hz bandwidth. request an acknowledgement message from the base transceiver
upon reception of packets that pass the error checking algorithm.
3.2. Wireless sensor network Data packets that fail to receive acknowledgements are placed in a
transmission queue for retransmission during a second scheduled
A wireless sensor is comprised of a traditional sensor and window. Packets are removed from the queue only upon successful
appropriate signal conditioning integrated with a transceiver unit receipt of their respective acknowledgement. The optimized
for the digital conversion and transmission of the raw signal from network transmission protocol coupled with high-speed base node
a remote measurement location to a central acquisition station communication transfers between the chip transceiver and the
without cabling. The wireless sensor network deployed utilized a host microcomputer are directly responsible for alleviating issues
commercial transceiver platform (MoteIV Tmote Sky) incorporated with data delivery, reduced sampling rates, and limited remote
with a custom software, signal conditioning hardware, and acquisition channels that have typically plagued wireless sensor
sensor design developed in-house to specifically address the networks.
requirements of bridge condition assessment through modal
analysis and load rating. The commercial transceiver platform 4. Test results
features a 4 MHz microcontroller interfaced with a 2.4 GHz chip
transceiver with a maximum data rate of 250 kbps. The device 4.1. Wireless sensor network performance
is compliant with U.S. and Canadian radio frequency regulations
and is certified by the FCC and Industry Canada for unlicensed use The average data success rate across all of the sensor nodes
in either country. A complete description of the wireless sensor over ten 3 minute test cycles was 99.91%, with 184 of the 200
network hardware and software as well as results of laboratory time histories reported with 100% data delivery success. The
validation can be found in Whelan and Janoyan [12]. minimum data success rate over these tests was 98.0%, which
The Wireless Sensor Solution (WSS) is a multi-functional sys- corresponds to a loss of only 17 data packets of the 850 requested
tem developed in-house by the authors to address both dynamic for the sampling duration specified. The small loss of data has
response and strain monitoring for load rating of in-service high- been attributed to a software bug discovered in the portion of
way bridges. Independent circuits accommodate signal condition- the embedded software code responsible for transmitting any
ing hardware specific to vibration measurement of typical bridge residual packets in the transmission queue after completion of
spectra and strain measurement from commercial transducers. The sampling. While correction of the software will likely improve
accelerometer conditioning section provides a two-channel inter-
the data success rates, the current level of data recovery is
face featuring a low-noise 3 V power supply, 14-stage digitally
generally more than sufficient for structural health monitoring
programmable signal amplification, auto-offset nulling, and a 5th
as the system identification analysis suffered from no noticeable
order analog low pass filter. A complementary signal conditioning
adverse distortion. For the network size and sampling rate of the
circuit for differential sensors, such as strain transducers, offers
deployment in this study, the radio protocol during active sampling
an application-specific integrated circuit (ASIC) tailored to high-
resulted in a transmission overhead in the range of 97 kpbs to
resolution measurements of resistive-bridge sensors. The chip fea-
126 kbps depending on the packet success and retransmission
tures programmable gain and offset of the differential signal, a
rates. The degree of transmission reliability attained at the high
15-bit charge-balancing ADC, a temperature interface with a dig-
data throughput rate prescribed in this testing demonstrates that
ital correction algorithm for compensating strain measurements,
wireless sensor networks are currently capable of performing
and nonvolatile memory for storing configuration settings during
large-scale, real-time structural health monitoring.
power cycling. Hardware shutdown of the individual signal condi-
tioning circuits as well as the transceiver provides for power con-
servation during periods of inactivity (Fig. 3). 4.2. Dynamic bridge response
Custom embedded software applications were written for
the sensor nodes to accommodate in-network task triggering A single-span integral abutment bridge provides an ideal
and large-scale concurrent node operation. Unique features of platform for testing the performance of a structural health
the software include digital filtering of sensor data using the monitoring system as the high stiffness of the bridge results in a
hardware multiplier, in-network remote sensor configuration, demanding measurement scenario in that structural vibrations are
control of component power supplies, and robust communication. of very low amplitude. Throughout the duration of testing, peak
An optimized radio transmission protocol that coordinates the accelerations across the structure ranged from less than 2 mg to
sampling and transmission software operations with node-based only nearly 10 mg. Despite this low excitation, the amplified sensor
scheduling is implemented in favor of the TinyOS radio stack signals produced clear time-history representations of the traffic
typically used for this transceiver platform, which relies on loading (Fig. 4) as well as distinct peaks in the frequency spectra.
concurrent processing of these tasks. Embedded software was The WSS signal conditioning hardware provides an extensive range
also written for the microcontroller at the host transceiver; of programmable gain amplification and therefore can adapt to the
a high-speed transparent bridge between the chip transceiver measurement range of bridges of various span lengths and support
and the virtual COM port operating at 262144 baud across conditions. Likewise, it can be reasonably assumed that the system
the USB bus enables significantly higher data throughput. A will even provide improved measurement performance on longer-
user-friendly LabVIEW software application controls bidirectional span structures with larger amplitude vibrations, as the signal-to-
communication through the base transceiver, configures the noise ratio will be increased.
sensor nodes remotely, displays real-time sensor data, and logs A particular concern with wireless sensor networks is that in
time histories to the hard disk for subsequent analysis. using localized self-contained data acquisition nodes distributed
The radio transmission protocol developed enabled nearly across a structure, there is no longer a shared sample clock
100% data delivery across the entire twenty node network. by which time synchronization of measurements can be strictly
2228 M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235

Fig. 3. Wireless sensor solution (WSS) hardware with MEMS accelerometer and BDI strain transducer.

Fig. 4. Typical acceleration time histories.

enforced. Conversely, all nodes operate with their own clocks


that will have unique offset and drift characteristics affecting
the relative timing of tasks among nodes in the network. Unlike
many wireless sensor networks that use the main microcontroller
oscillator to provide sample timing, the WSS system relies on
an accurate 32.768 kHz crystal oscillator to provide a precise,
stable clock to each of the nodes. Synchronized initiation of
sampling is invoked by a single command broadcast to all nodes
in the network. Since the radio packet is an electromagnetic
wave, therefore travelling at the speed of light, the difference in
reception time amongst the nodes in the network will be on the
order of nanoseconds. Consequently, the accuracy and stability
of the crystal oscillators dictates the time synchronization of
the samples relative to each node throughout time. Investigation
of the accelerations measured across the network during traffic
events reveals that the wireless sensors maintain phase amongst Fig. 5. Time window overlay of accelerations during a traffic event.
themselves, which is imperative for accurate operational modal
analysis (Fig. 5). This thereby substantiates the assumption that an forced vibration to the bridge would require at least a partial
accurate crystal oscillator can provide time synchronization of the closure of traffic and does not lead itself to long-term continuous
network for at least the sampling duration utilized. or autonomous structural health monitoring. Furthermore, the
presence of any nearby vibration sources, such as machinery or
4.3. Operational modal analysis traffic, and the effect of dynamic environmental loads, such as
wind, geological, and hydrodynamic forces, are additional system
Analysis of in-service dynamic measurements of highway inputs that are not accounted for in the measurement of the
bridges challenges conventional modal analysis techniques as controlled excitation input. Consequently, extraction of system
it is often impractical to measure the input excitation. The modal parameters of civil structures from ambient excitation
use of impulse excitation (drop-weight) or a shaker to provide has recently emerged as a widely sought-after approach for
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2229

estimate, the operational deflection shapes are generally similar


to the mode shapes. However, for structures with closely spaced
modes, the estimated mode shapes will be a combination of
multiple modes and therefore may not produce accurate response
estimation. Fortunately, the structure tested has adequately spaced
natural frequencies with sufficient damping to permit clear
extraction of many of the dominant modes within the measured
bandwidth. However, the traffic patterns captured in the field
testing either did not adequately excite several additional modes
identified in the subsequent SSI analysis to permit identification
in the frequency spectra and/or these mode were damped
sufficiently to prohibit visual identification. Furthermore, while
the second longitudinal bending mode (4th mode) was identified
and estimated, the peak was relatively small in the spectrum and,
as evident in the mode shape estimate, the low signal-to-noise
ratio resulted in a relatively coarse approximation.
Fig. 6. Average normalized power spectral density. It should be noted that although the 6th and 8th modes
identified closely resemble each other, they are in fact distinct
in-service condition assessment. Traditional system identification modes as evidenced by several indicators. The reason that they
methods requiring both input and output measurements, such as appear so similar is that the 6th mode is a largely girder-based
using the frequency response function (FRF), cannot be applied mode shape, while the 8th mode is characterized by deck motion
to data yielded from ambient excitation, as there is no measure that produces similar consequent girder deflection. When the
of the system input. Recent work in the development of ambient frequency spectra is examined, these modes produce distinct
vibration methods has produced several approaches for output- and well-excited poles, which is anticipated as the width and
only system identification to enable modal parameter extraction lane loading associated with vehicular traffic is conducive to
from ambient time histories [13,14]. In this study, output- excitation of these modes. The presence of the same mode shapes
only system identification from the accelerometer time histories in subsequent stochastic subspace identification furthers this
was performed using the Modal Analysis on Civil Engineering rationale; close examination of the finite element analysis results
Constructions (MACEC) software package [15]. This software fully illustrates this initially subtle difference in the actual mode
analysis package permits the use of both the classical Fourier- shapes.
based Frequency Domain Decomposition (FDD) with peak peaking While the Fourier-based peak-picking method analyzes the
as well as the stochastic subspace identification (SSI) method for frequency-spectrum to extract natural frequencies and mode
deriving mode shapes, natural frequencies, and damping ratios. shapes, stochastic subspace identification techniques use a time-
During implementation of FDD modal analysis, a 4096- domain driven approach to solving for modal parameters from a
point average normalized power spectral density (ANPSD) was discrete-time stochastic state–space model. The derivation of this
computed using the time histories from all test sequences method is outside the scope of this paper; full presentation of
and sensor locations (Fig. 6). Natural frequencies were selected the state–space equations and model development can be found
from the modal peaks present in the power spectrum and the elsewhere [14], [16]. In short, the method uses a state–space model
corresponding mode shapes were derived for each test sequence. developed from the linear, second-order differential equation for
In this approach, discrete mode shape data are determined using a multi-degree of freedom spring, mass, and damper system
the relative magnitude and phase angle of the spectral peaks with input forces. To account for inherent sensor noise and
at each sensor location, corresponding to the eigenfrequency, fixed-width computational effects, the state–space model includes
with respect to a specified reference sensor. Since multiple time stochastic terms to represent measurement and process noise that
histories were available for analysis, the derived mode shapes are assumed to have zero-mean, white noise characteristics. The
were then normalized and averaged to provide the final mode MEMS accelerometers utilized in this study do in fact exhibit
shape estimates (Fig. 7). Averaging reduces the effects of noise a white noise spectrum according to electrical specifications, so
and improves the resolution of the mode shapes. Due to the mass spurious poles should not arise in the model as a result of violations
of passing vehicles, the mode shapes and natural frequencies of of this assumption. Since the system inputs are unknown and
the unloaded bridge will be slightly affected by traffic loads and immeasurable in the case of ambient excitation, the input terms
patterns. When using traffic excitation, averaging mode shapes are then implicitly modeled within the noise terms resulting in
from a large number of tests also alleviates the effect of traffic load a purely stochastic system. It should be noted that by modeling
bias on one lane of the bridge, which is more probable from the the unknown excitation forces with the noise terms, the inputs
results of only a single test. are assumed to also exhibit zero-mean, white noise characteristics.
Application of the FDD method on the experimental data was Violation of this assumption may produce spurious poles in the
found to produce smooth mode shapes that correlate well with the state–space model that are not inherent to the dynamics of the
FEA analysis. However, this method relies on sufficient excitation structure but arise from spectral bias within the excitation force.
of the eigenfrequencies to permit identification of each modal Once the system inputs to the state–space model are reduced
peak in the average normalized power spectrum. Additionally, the solely to the stochastic terms, numerical methods can be used to
mode must be sufficiently under-damped such that resonance at solve for the state–space matrices from the measurement data in
the natural frequency is represented by a distinct peak. In fact, order to produce a mathematical description of the system from
the frequency domain peak-picking method actually produces which all the modal parameters, except the mode scaling factor,
operational deflection shapes rather than mode shapes in that can be determined.
the shape constructed from the spectral data is the naturally Although the mathematical formulation of the SSI method and
weighted combination of all mode shapes that would arise if subsequent numerical solution is rather rigorous, especially to civil
the structure was excited by a pure harmonic at the selected engineers unfamiliar with systems and control modeling, the ap-
natural frequency [16]. Since only spectral content near the natural plication of the method is facilitated through the MACEC soft-
frequency contribute noticeably to the constructed mode shape ware environment, which requires only a basic understanding of
2230 M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235

Fig. 7. Operational deflection shapes extracted by Frequency Domain Decomposition (FDD) (numbering corresponds to mode as identified in subsequent FEA analysis).

Fig. 8. Principal angles between subspaces and stabilization plot for typical SSI analysis (⊕- stable pole, ·- pole with partial pass of stability criteria).

state–space models and general system identification methodol- which correlates to twice the number of system poles, or eigenfre-
ogy. Following calculation of the system model from the measure- quencies, captured in the measurement bandwidth. Unfortunately,
ment data, the principal angles between subspaces can be plotted real-world data rarely produces a single, distinct gap, as evidenced
to provide a means of estimating the model order (Fig. 8(a)). A gap in the results for the Wright Road Bridge analysis in which gaps in-
in the subsequent principal angles indicates the model order [15], dicate several possible model orders ranging from as low as six and
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2231

yields a greater number of modes with higher quality shape


estimation than the FDD technique for output-only modal analysis.
Given the insignificant number of spurious poles in the state–space
model, particularly in the portion of the spectrum above 8 Hz
where the dynamic structural response occurs, the signal-to-noise
ratio is low enough to permit reliable extraction of these twelve
modes. It will be shown in the subsequent section that by over-
specifying the model order, as suggested in the MACEC manual,
the remaining two structural modes can be extracted from the in-
service measurements.

5. Numerical model of test structure

Fig. 9. Average spectral density of computed SSI state–space model (order 44, poles A solid model of the bridge tested consisting of the steel
indicated with star).
superstructure, concrete slab, abutments, and railings was drafted
in AutoCAD 2004 from the as-built Department of Highways
as high as 44. Given this all too common scenario, over-specifying drawings (Fig. 11). Finite element analysis of the model was
the model order is recommended [15] and the suggested model or- performed using the FEMPRO software package by ALGOR
der can be determined from the number of stable poles identified in Incorporated. Natural frequency (Modal) with Load Stiffening
the stabilization diagram (Fig. 8(b)). At this point, subjectivity is re- analysis was performed using three-dimensional solids consisting
quired to select the poles likely to be a result of structural dynamics of brick elements. A mesh size of 20.32 cm (8 in.) was specified
rather than spurious poles resulting from the numerical process. in the auto-mesh generation, which resulted in 33768 total solid
The previously outlined approach to extracting modal param- elements across the model. Abutment and slab properties were
eters from a SSI model, as suggested by the MACEC manual, can specified by assigning the material type to medium-strength
be treated as a relatively simple means of quickly extracting mode concrete as defined by the built-in library, while the steel
shapes, natural frequencies, and damping ratio estimates from superstructure and railings were assigned the material properties
experimental data. However, through experience with applica- of A588 and A500 steel, respectively.
tion of output-only system identification to in-service dynamic re- Modeling the behavior of an integral abutment bridge is
sponse measurements from highway bridges, the authors of this substantially more complex than developing a similar model for
paper highly recommend developing a single state–space model a deck supported by bearings, as the soil–structure interaction on
from which all modal parameters are estimated. In this manner, the abutments and piles has a significant influence on the dynamics
subjectivity in identification of actual structural poles versus spu- of the superstructure [17]. Given that the analysis was performed
rious poles can be greatly reduced and the derived modal parame- simply to provide a plausibility check with the measured response,
ters will reflect a more consistent estimate of the system response some assumptions and modeling simplifications were made in
less affected by spurious poles. The appropriate model order can be providing boundary constraints and loads on the surfaces of the
coarsely predicted by doubling the number of spectral peaks evi- abutments. Lateral soil pressure on the abutments and wingwalls
dent in the average power spectrum and then more accurately esti- was provided using an assumed linear pressure distribution, a
mated by finding a gap in the principal angle plot (Fig. 8(a)) greater 1.922 Mg/m3 (120 lb/ft3 ) unit weight of the backfill soil, and a
than the coarse model order prediction. This process can be aided coefficient of lateral earth pressure of 0.5. This soil pressure results
by comparing the spectral density estimate of state–space model in a slight decrease in natural frequencies, due to the longitudinal
to the power spectrum of the measured data. For the case study compressive force applied on the deck. While nonlinear spring
presented, a model order 44 was determined to be most appropri- elements are often used to model soil–structure interaction, the
ate; the discontinuity in the principal angle plot was found to be FEA software allowed only for linear elastic spring elements in the
the last significant gap in the profile. The average spectral density natural frequency with load stiffening analysis. Consequently, the
of the 44th order state–space model (Fig. 9) correlates well with National Cooperative Highway Research Program (NCHRP) curves
spectral content of the ANPSD computed from the measurement for earth pressure coefficient versus relative wall displacement, as
data (Fig. 6) and the majority of poles have been found to arise from presented by Civjan [18] were approximated with a linear fit to
the structural response. Plotting the poles over the average spec- the initial slope of passive pressure development. Since the backfill
tral density of the state–space model provides an efficient means of soil properties are unknown, the linear slope approximations
identifying which poles are spurious and which relate to the struc- for loose and dense sand were averaged. Translational spring
tural dynamics of the structure. Structural poles, i.e. the natural fre- elements were placed on vertices across the abutment in both
quencies of the bridge, coincide with distinct peaks in the average the longitudinal and transverse directions of the bridge deck. The
spectral density, whereas the spurious poles generally fall between elastic modulus of each spring element was assigned with a linear
peaks for a model of appropriate order. profile increasing with depth, as determined by the approximation
Modal parameters were then extracted from the state–space to the NCHRP curve. Translational stiffness was applied to the base
model to estimate the natural frequencies and damping ratios of of the abutment to represent bearing pressure and frictional forces
the in-service bridge. Twelve non-spurious poles were identified in along this surface. A moderate 350.4 kN/m (31.25 lb/in/in2 ) spring
the model permitting successful extraction of twelve mode shapes stiffness was assigned to the lateral springs on these vertices;
from the measurement data, thereby demonstrating significant vertical springs were applied with 14.01 MN/m (1250 lb/in/in2 )
improvement over the application of FDD system identification stiffness. The contribution from support piles was modeled with
(Fig. 10). In addition to identifying mode shapes that are less additional translational springs at the vertices corresponding to
pronounced in the frequency spectrum, the SSI analysis resulted pile locations. Although the pile axis orientations differed at each
in smoother mode shapes, alleviating signal-to-noise issues for abutment, the translational springs applied to represent the piles
poorly excited modes. Consequently, although the SSI technique assumed equal stiffness in both lateral directions.
requires substantially more understanding of system identification The superstructure-dominant mode shapes, i.e. those measur-
and state–space modeling, is computationally more complex, and able through the accelerometer placement on the girders, esti-
necessitates a higher degree of subjectivity, the process effectively mated within the measurement bandwidth through the FEA are
2232 M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235

Fig. 10. Operational mode shapes present in SSI model of 44th order (3D surface plot and plan view with magnitude shading; Numbering corresponds to mode as identified
in the subsequent FEA analysis).

Table 1
Comparison of experimental and FEA natural frequency estimates.
Mode shape Classical peak-picking method (Hz) Stochastic subspace identification (SSI) (Hz) Damping ratio (%) from SSI Finite element analysis (Hz)

1 9.5 9.58 1.60 9.18


2 11.09 11.14 1.03 11.26
3 17.66 17.74 1.10 19.07
4 22.44 18.71 17.15 22.33
5 – 26.92 7.59 25.69
6 29.97 30.03 0.82 34.65
7 31.56 31.47 0.26 35.61
8 33.78 33.89 0.97 57.87
9 35.38 35.21 1.23 35.74
10 46.88 46.98 0.59 49.14
11 – 47.96a 10.14 50.34
12 – 51.73a 9.76 57.13
13 53.09 54.04 2.54 53.19
14 – 62.01 / 65.99 0.86 69.65
a
Not in 44th order model, identified through over-specifying model order.

presented in Fig. 12. While additional mode shapes were identi- the natural frequencies between the FEA analysis and the exper-
fied in FEA, these mode shapes were generally either dominated imentally measured modes, particularly at the higher end of the
by slab or wingwall movement rather than motion at the gird- spectrum. Given the bending orders of the modes, it is natural to as-
ers or the result of rigid body translation and rotation on the sume that, for instance, the third order longitudinal bending mode
boundary springs. Comparison of the analytical modal parame- with first order bending in the lateral direction will occur at a lower
ters with those identified in the experimental system identification frequency than the third order longitudinal bending mode with
techniques yielded good correlation in terms of both visual second order bending in the lateral direction despite the discrep-
comparisons of mode shapes and relative estimation of natural ancy in the results of the FEA analysis. These differences can be at-
frequencies (Table 1). There are some differences in the order of tributed to the assumptions, approximations, and simplifications
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2233

Fig. 11. Finite element model: (a) mesh, (b) loads and boundary constraints.

made in the finite element model. Consequently, the mode shapes Examining the spectral density of the state space model, it becomes
have been sequentially numbered according to the proper order as apparent that these modes suffer from a lack of excitation as
measured experimentally. well as significant damping. Furthermore, after introducing the
It should be noted that the four highest order modes have additional degrees of freedom necessary for the corresponding
natural frequencies that reside in the portion of the bandwidth poles to arise in the model, there are significantly more spurious
with potential aliasing. As a result, these natural frequency poles introduced. These spurious poles are not intrinsic to the
estimates may be incorrectly aliased from the 64–78 Hz frequency structural response, but arise from noise and excitation violations
range. However, due to the model order associated with the of the zero-mean, white noise assumptions. Consequently, the
bending patterns, it is likely that all of the modes, aside from the over-specified model is not recommended for implementation in
last, can be assumed to be correctly associated with the 50–64 Hz any further analysis, such as forward prediction with a Kalman
band, as the last mode is of the highest bending order. Furthermore, filter. It is recommended to use the lower order model even though
signal attenuation in the aliased region for these modes is also it does not include two of the fourteen modes known to be present
significant enough that identification of an aliased peak in the in the bandwidth of interest.
spectrum would be highly unlikely. Consequently, only the last
mode has been identified with two possible natural frequency 6. Conclusion and discussion
estimates for the experimental analysis. While some discrepancies
exists among the analytical and experimentally derived modal To field-test the performance of a wireless sensor network
parameters, the complexity of modeling an integral abutment for dynamic response assessment of in-service highway bridges,
bridge coupled with modeling assumptions could accounts for a single-span integral abutment bridge has been investigated
these slight differences as well as any inconsistencies between through operational modal analysis using ambient vibration
the bridge design and actual construction tolerances and material testing. As a consequence of the experimental program, the
properties. Wireless Sensor Solution (WSS) platform for structural health
Given that the FEA identified two additional superstructure- monitoring developed at Clarkson University has demonstrated
dominant modes within the measured bandwidth, stochastic the capability to replace cable-based instrumentation for the
subspace identification was revisited to employ a higher model majority of in-service condition assessment routines. In addition,
order to extract these missing modes (Fig. 13). At a model order of the alleviation of obstacles associated with cabling, such as the
76, the remaining two modes became apparent in the state–space associated installation time and cost, introduces the potential to
model, though the shapes are only approximate due to the low monitor the performance of an increased number of bridges in a
excitation and large damping ratios associated with these poles. more condensed time frame.
2234 M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235

Fig. 12. Finite element analysis of mode shape development.

Fig. 13. Remaining experimental mode shapes extracted through over-specification of model order in stochastic subspace identification (respective poles identified over
average spectral density of 76th order model).

The primary objective of the study was to evaluate the perfor- for well-excited modes in good agreement with finite element
mance of the wireless sensor network and local data acquisition analysis. However, the use of the SSI technique permitted the ex-
hardware in a typical field setup outside of the laboratory environ- traction of an additional four modes from the time histories in
ment. Real-time streaming of 40-channels of measurement data addition to damping ratio estimates for all vibration modes. Fur-
sampled at an effective rate of 128 Hz per sensor for ten test du- thermore, the relative quality of the mode shapes derived through
rations each exceeding three minutes was successfully achieved SSI was deemed to be higher than obtained from FDD and is likely a
while maintaining nearly 100% data delivery across the network. result of the inclusion of stochastic noise components in the math-
Additionally, the signal conditioning and data acquisition inter- ematical formulation of the SSI state–space model. In general, the
face provided an in-network reconfigurable platform that enabled use of SSI techniques to estimate modal parameters from output-
sufficient resolution of the low amplitude vibration experienced only experimental data has been found to be preferable to the FDD
to extract fourteen mode shapes of the relatively stiff single-span method despite the increased computational effort and subjectiv-
bridge using low-cost MEMS accelerometers. These results present ity required to identify system poles. Lastly, the authors present
a significant breakthrough in the use of wireless sensor networks an approach for estimating modal parameters using a single or-
for structural health monitoring. By essentially replicating a typi- der state–space model developed through the SSI system identifi-
cal cable-based dynamic test routine in terms of not only sampling cation, rather than selecting poles from a stabilization diagram.
rates, number of deployed sensors, and test duration but also in the
quality and breadth of extracted modal parameters, the developed
Acknowledgements
wireless sensing platform has emerged as both a feasible and com-
parable alternative to wired instrumentation in structural health
monitoring and in-service condition assessment. This research has been funded by the New York State Energy
In analyzing the ambient vibrations measured by the wireless Research and Development Authority (NYSERDA), in collaboration
sensor nodes, a complementary study was performed in which with the St. Lawrence Highway Department, and the New York
the effectiveness of applying two output-only system identifica- State Department of Transportation (NYSDOT). The authors would
tion methods to real-world measurements was evaluated. The also like to acknowledge the assistance of Kevin Cross and Dan
frequency domain decomposition technique was compared to Nyanjom during the field deployment and Michael Fuchs with
stochastic subspace identification to contrast the ability of the two system development. Any opinions, findings, and conclusions or
methods to extract modal parameters. Overall, FDD was capable recommendations expressed in this paper are those of the authors
of constructing mode shapes and estimating natural frequencies and do not reflect the views of the agencies.
M.J. Whelan et al. / Engineering Structures 31 (2009) 2224–2235 2235

References [9] Paek J, Jang OGK-Y, Nishimura D, Govindan R, Caffrey J, Wahbeh M, Masri S. A
programmable wireless sensing system for structural monitoring. 4th world
[1] Wardhana K, Hadipriono FC. Analysis of recent bridge failures in the United conference on structural control and monitoring, San Diego (CA); 2006.
States. J Performance Construct Fac 2003;17(3):144–50. [10] Lynch JP, Wang Y, Loh KJ, Yi J-H, Yun C-B. Performance monitoring of the
[2] Farrar CR, Baker WE, Bell TM, Cone KM, Darling TW, Duffey TA, Eklund A, Geumdang bridge using a dense network of high-resolution wireless sensors.
Migliori A. Dynamic Characterization and Damage Detection in the I-40 Bridge Smart Mater Struct 2006;15(6):1561–75.
over the Rio Grande. Los Alamos National Laboratory, Technical Report LA- [11] Wenzel H, Pichler D. Ambient vibration monitoring. West Sussex, England:
12767-MS; 1994. John Wiley & Sons Ltd; 2005.
[3] Alampalli S, Fu G, Dillon EW. Signal versus noise in damage detection by [12] Whelan MJ, Janoyan KD. Design of a robust, high-rate wireless sensor network
experimental modal analysis. J Struct Eng 1997;123(2):237–45. for static and dynamic structural health monitoring. J. Intell. Mater. Syst.
[4] Maeck J, Peeters B, De Roeck GD. Damage identification on the Z24 bridge using Struct.. 2009 (in press) doi:10.1177/1045389X08098768.
vibration monitoring. Smart Mater Struct 2003;10:512–7. [13] Ventura CE, Felber AJ, Stiemer SF. Experimental investigation of dynamics of
[5] Huth O, Feltrin G, Maeck J, Kilic N, Motavalli M. Damage identification using Queensborough bridge. J Performance Construct Fac 1995;9(2):146–55.
modal data: Experiences on a prestressed concrete bridge. J Struct Eng 2005; [14] Peeters B, De Roeck G. Reference-based stochastic subspace identification for
131(12):1898–910. output-only modal analysis. Mech Syst Signal Process 1999;13(6):855–78.
[6] Wipf TJ, Phares BM, Klaiber FW, Wood DL, Mellingen E, Samuelson A. [15] Van den Branden B, Peeters B, De Roeck GD. Introduction to MACEC v2.0:
Development of bridge load testing process for load evaluation, Iowa DOT Modal analysis on civil engineering constructions user guide and case studies.
Project Report TR-445; 2003. Katholieke Universiteit Leuven; 1999. p. 47.
[7] Lynch JP, Loh KJ. A summary review of wireless sensors and sensor [16] Peeters B. System identification and damage detection in civil engineering.
networks for structural health monitoring. Shock Vibration Digest 2006;38(2): Ph.D. dissertation, Katholieke Universiteit Leuven. 2000.
91–128. [17] Fennema JL, Laman JA, Linzell DG. Predicted and measured response of an
[8] Pakzad SN, Kim S, Fenves GL, Glaser SD, Culler DE, Demmel JW. Multi-purpose integral abutment bridge. J Bridge Eng 2005;10(6):666–77.
wireless accelerometer for civil infrastructure monitoring. In: Proceedings [18] Civjan SA, Bonczar C, Breña SF, DeJong J, Crovo D. Integral abutment bridge
of the 5th international workshop on structural health monitoring, Stanford behavior: Parametric analysis of a Massachusetts bridge. J Bridge Eng 2007;
(CA); 2005. 12(1):64–71.

You might also like