Open access
Technical Papers
Feb 18, 2019

Remote Bridge Monitoring Using Infrasound

Publication: Journal of Bridge Engineering
Volume 24, Issue 5

Abstract

As transportation infrastructure continues to age, new methods of noncontact sensing should be evaluated and, if found suitable, used for bridge monitoring and structural health assessment. This study highlighted the use of infrasound monitoring, a geophysical technique utilizing acoustics below 20 Hz, as one possible solution for noncontact, nonline-of-sight bridge health monitoring. The study focused on the technique of infrasound for infrastructure monitoring with a detailed case study involving a steel, two-girder bridge in northern California. Infrasound was used to detect natural modes of the structure from a distance of 2.6 km. The frequencies detected infrasonically were validated with data collected by on-structure accelerometers. The noncontact nature of this structural assessment approach has potential to supplement traditional structural assessment techniques as affordable, remote, persistent monitoring of transportation infrastructure. Implications for use of this technology were also discussed alongside specific applications for scour monitoring and postdisaster assessment.

Introduction and Background

As the transportation infrastructure of the United States ages, the cost of inspection and monitoring is increasing, necessitating consideration of new and complementary methods for bridge monitoring and structural health assessment. The current inspection process is labor intensive and requires full access to the structure. Furthermore, current practice most often captures discrete points in time rather than providing persistent monitoring. By law, all bridges within the US National Bridge Inventory (NBI) must be inspected every 2 years. If problems are identified, the inspection regularity must be increased (FHWA 2004). Even with an increased inspection cycle, there are still long time periods between inspections. With more than 611,000 bridges currently in service in the United States (FHWA 2017), introduction of complimentary processes to the current hands-on inspection procedures in the form of noncontact, remote sensing might allow for better prioritization of limited resources as well as continuous monitoring of structures with deficiencies deemed candidates for persistent monitoring.
In the last decade, various methods of remote sensing have been investigated for application to structural health monitoring, including light detection and ranging (LiDAR), photogrammetry, infrared thermography, radio detection and ranging (RADAR), multispectral satellite imagery, synthetic aperture radar (SAR), and image analysis methods, including digital image correlation (DIC), ground penetrating radar (GPR), and remote acoustics (Chen et al. 2011; Vaghefi et al. 2012; Fukuda et al. 2013; Gentile and Cabboi 2015; Harris et al. 2016). Vaghefi et al. (2012) and Harris et al. (2016) compared several of the previously listed methods to investigate bridge assessment and monitoring performance. Results indicated that the techniques worked most effectively when used in conjunction with one another (Harris et al. 2016), whereas specific groupings of sensing technologies were most efficient at identification of defects in specific locations (Vaghefi et al. 2012). Of these technologies, not all can be considered noncontact sensing, meaning they still need direct access to the structure, thus reducing their applicability for remote sensing. The remainder of this paper is devoted to presenting infrasound analysis as a method for noncontact, nonline-of-sight, remote bridge monitoring.

Infrasound for Infrastructure Monitoring

Engineers are very comfortable with the understanding that structures resonate, but what is not often considered is that the air surrounding these structures is a fluid that can propagate pressure variations caused by structural vibration. The perturbations caused by the resonance of these structures create sound in the infrasonic range that propagates distances of tens of kilometers from the source structure while retaining critical frequency information about the structure. Essentially, infrasound monitoring allows for remote acoustic detection of the natural frequencies of the structure being monitored.
Natural frequencies are the lowest natural modes of vibration for a structure and are unique to each as a function of the mass and stiffness of that structure (Chopra 2012). These modes may shift slightly due to thermal or other effects through the course of a day or seasonally; however, a sudden and significant shift in these natural frequencies, particularly the lowest modes, is indicative of a structural change that warrants inspection (Salawu 1997). Two forms of excitation can be used to excite these modes: ambient excitation (such as wind, traffic, or seismicity) and forced excitation (such as mass shaker or modal hammer) (Farrar et al. 1999; Conte et al. 2008; Cunha et al. 2013). Conte et al. (2008) indicated that ambient excitation may be more efficient at exciting the lowest natural frequencies, leaving no need for forced excitation.
Among the sources of ambient excitation, traffic is unique in that it tends to excite the higher-order modes of vibration. Thus, the question arises concerning how robust infrasound monitoring can be when traffic is present. There are two considerations for frequencies related to traffic. The first is the transient nature of traffic sources, and the second is the frequency range of traffic sources. Because the signals from traffic are transient, they can be decoupled from the continuous source signal produced by a bridge. The frequencies for traffic sources, both those related to the traffic itself and those in the bridge excited by the traffic, tend to be higher (Conte et al. 2008). These higher frequencies are less efficient at propagation than the lower frequencies due to signal attenuation. At distances of a few kilometers, signal attenuation is negligible. However, in cases such as McKenna et al. (2009c), in which the far-field sites of 20 and 27 km were able to detect the bridge of interest, attenuation of the higher frequencies does become a concern, with the lower frequencies, more specifically those below 5 Hz, propagating more efficiently (Sorrells et al. 1997; Bass et al. 2006).
The strength of the signal from actual vehicular traffic is not strong enough to propagate to distances at which infrasound arrays would be deployed, from 2 to 25 km. Likewise, the higher modes of the bridge excited by traffic are not primarily bending modes, meaning they will move less air than the lowest modes, which also limits propagation distance. In essence, infrasound seeks to evaluate lower-frequency responses at longer distances as opposed to higher transient responses (e.g., traffic) at shorter distances. These factors make infrasound monitoring a unique technology to remotely observe a structure for global condition changes with no need for contact with or line of sight to the structure. Most other technologies require at least one of these items.
Infrasound is defined as acoustics below 20 Hz (Bedard and Georges 2000; Evers 2005; Christie and Campus 2010). The long wavelengths associated with these low frequencies allow for propagation of a signal with little attenuation, meaning the frequency content of the signal is preserved. Even though infrasound has long been used by geophysicists and the nuclear monitoring community as part of the Comprehensive Nuclear-Test-Ban Treaty (CTBT), the use of infrasound in relation to infrastructure monitoring is relatively new. The first recorded case of bridge monitoring using infrasound was reported by Donn et al. (1974) related to the Tappan Zee Bridge in New York. An intermittent signal was detected in bursts for several hours over a period of months by the geophysical group at the Lamont-Doherty Geological Observatory in Palisades, New York. Further investigation and data processing yielded a back azimuth, the direction from which the processed signal originated, that triangulated to the Tappan Zee Bridge, with frequencies observed by infrasound being corroborated with on-structure geophones (Donn et al. 1974).
At the time of the experiment by Donn et al. (1974), there was little overlap between the geophysical and structural engineering communities, and advancement of infrasound monitoring of civil infrastructure remained stagnant until an experiment series performed by the US Army Engineer Research and Development Center (ERDC) beginning in 2006. A joint team of geophysicists and structural engineers investigated a steel, through-truss railroad bridge over the Little Piney River in Ft. Leonard Wood, Missouri, in a combined effort that included traditional engineering methods of bridge assessment complemented by infrasound monitoring (Diaz-Alvarez et al. 2009; McKenna et al. 2009a, b, c). Natural frequencies of the structure were predicted by a finite-element model calibrated through data collected during a load test of the bridge. This model was also used to determine if the identified frequencies fell within the infrasound passband. In addition, three infrasound arrays were set up to monitor the bridge: one at the bridge itself (<1 km), and two more at distances of approximately 20 and 27 km. Analysis of the data collected from the two far-field infrasound arrays identified a continuous-wave packetized signal with a frequency content that matched the natural frequencies observed with on-structure instrumentation, and frequency-wave number analysis yielded back azimuths that aligned with the structure.

Experimental Series

The bridge of interest for the current experimental series, denoted as Br 18–0009 in the NBI, is located in northern California and carries Route 20 between Marysville and Yuba City over the Feather River. The scour-critical, 24-span, steel, two-girder bridge with a longest span of 49 m was monitored by a network of three arrays composed of infrasound sensors arranged at distances ranging from 2.6 to 24 km. In the first experiment, infrasound data were collected for analysis in this study. In the second experiment, the primary focus was on-structure instrumentation (accelerometers) to validate the collected infrasound data.

Experiment 1

Infrasound sensors are commonly deployed in networked, multisensor arrays for localization and characterization of signals; thus, multiple arrays are required for optimal data. For this experiment, three arrays were deployed with sites identified through considerations for space availability, security, wind protection, topography, distance from source to receiver, and prevailing meteorological conditions (Simpson et al., Forthcoming). Arrays were located at the Sutter County Airport, the Marysville City Cemetery, and Beale AFB at distances of 3.0, 2.6, and 24 km, respectively. Each of these array locations was separated from the bridge of interest by urban terrain, industry, and/or significant topographical features, such as levee systems. For analysis, these arrays were denoted as FRA for the Sutter County Airport array, FRB for the Beale AFB array, and FRC for Marysville City Cemetery array. Fig. 1 presents the location of each array in relation to the bridge as well as a layout of each array.
Fig. 1. Array layout for Experiment 1. Array locations at the Sutter County Airport (FRA) and Marysville City Cemetery (FRC) were also used in Experiment 2.
Data collection took place over 7 days from February 26 through March 4, 2014, with data sampled at 1,000 Hz, and the gain was set at unity for all infrasound sensors. Each of the five-element infrasound arrays consisted of five infrasound sensors with porous hoses attached to the four inlet pipes as passive wind filters, two digitizers, and associated cables, connectors, batteries, and solar panels. Some specifications for the equipment used are included here, but additional information can be found in the work by Simpson et al. (Forthcoming). The Inter-Mountain Laboratories (IML, Sheridan, Wyoming) Model ST infrasound sensors utilized in this experiment had a nominal signal frequency band of 2–30 Hz, but content could extend below 2 Hz with roll-off beginning at approximately 1 Hz. Data from these components were streamed to REF TEK (Plano, Texas) 130S-01 digitizers with an input preamplifier and digital antialias filters and a high-precision external Global Positioning System (GPS) receiver/clock for time synchronization between each digitizer through the course of the experiment. Arrays were deployed in a 60 × 60-m cross pattern with infrasound sensors at the cardinal points and one infrasound sensor at the center. The cross configuration gives omnidirectional detection capabilities while simplifying data processing.
Understanding atmospheric conditions is imperative to monitoring infrasound, because the atmosphere is a time-varying propagation medium with temperature and wind effects dominating propagation characteristics. During this experiment, radiosondes were launched to capture the atmospheric conditions. The radiosondes sampled the atmosphere once per second from balloon release at ground level to balloon bursting height at maximum altitude (balloon dependent). This deployment used 350-g balloons reaching an average altitude of approximately 25 km and a travel time of approximately 2 h. Data obtained by the radiosondes included temperature, pressure, humidity, wind direction and speed, altitude, and GPS coordinates. To minimize interruption to flight operations and training at Beale AFB, launches were limited to every 6 h, starting at 0600 on March 1 and continuing through March 2, 2014. Radiosonde launches occurred from the Sutter County Airport.
Two REF TEK 130-SMA strong motion accelerographs were also deployed at Br 18–0009 (one on the sidewalk and one directly below it at the base of a pier, with both pointing north) on March 3, 2014, with the intention of gathering data for 24 h to try to capture the natural frequencies of Br 18–0009. The intent was to use these frequency data as validation for the collected infrasound data. Unfortunately, the sensors were vandalized while deployed, rendering the data unusable. A second experiment was conducted to try to capture these on-structure data as well as collect additional infrasound data in the area. Contact with Caltrans indicated that conditions at the structure had not changed from the first experiment to the second, making comparison of infrasound data from the first experiment to the on-structure instrumentation results in the second experiment valid.

Experiment 2

The second experiment took place November 13–18, 2015, and focused primarily on capturing the fundamental modes of Br 18–0009 utilizing on-structure instrumentation. In addition, two of the three array locations from the first experiment were reoccupied. Initial data-analysis efforts yielded no detection of the low-energy signals at FRB during the initial times of interest (temperature inversions) identified in the collected radiosonde data; therefore, the decision was made not to instrument this location for the second experiment. The arrays at the Sutter County Airport (FRA) location and the Marysville City Cemetery (FRC) location utilized the same instrumentation as described for the first experiment. A single sensor was placed at the bridge during this experiment as well to try to collect a more direct infrasound measurement from the bridge. The Hyperion (Tupelo, Mississippi) IFS-3000 series infrasound sensor located at the bridge utilized a dome wind filter rather than porous hoses and has a response within 3-dB variation from 0.01 to 100 Hz. Data collection took place November 13–18, 2015. In addition to infrasound sensor arrays, accelerometers were placed on the main spans of the bridge from November 16–18 to validate frequencies detected through infrasonic monitoring.
Twelve Gulf Coast Data Concepts (Waveland, Mississippi) multifunction extended life accelerometer data logger –×2 units (operated in high-gain mode), such as the one presented in Fig. 2, were used for continuous collection of bridge vibration data in this experiment. These devices combine a battery-powered digitizer and a Class C microelectromechanical system (MEMS)-based accelerometer. The accelerometer is a three-axis device with a range of ±2 g, whereas the data recorder can digitize up to 512 S/s on each channel with 16 bits of resolution. Each digitizer had 32 GB of storage capability, giving ample storage for the 46-h time frame during which the units were deployed. For a detailed description of this sensor’s performance and its comparison with similar devices, the reader is referred to work by Evans et al. (2014). Due to a significant time-stamping error detected with these units at their highest sample rate (512 S/s), the data in this experiment were collected at a sample rate of approximately 255 S/s. The data logger was configured to sequentially record files in 15-min blocks in its high-resolution time-stamped data-collection mode. The time-stamp information was accurate to 0.1 ms and was used during postprocessing to resample the data to a uniform rate. Individual data-logger clocks were synchronized, and cross-correlation-based delay estimation was used to perform final time alignment of the data.
Fig. 2. (Color) Location of accelerometers on Spans 21 and 22 of Br 18–0009 with an example of a deployed sensor and its collocated backup sensor.
These sensors were deployed on Br 18–0009 in three locations at Span 21 and in three similar locations at Span 22. The use of switchable magnets on the sensor-mounting plates, as seen in Fig. 2, allowed for rapid and slip-free attachment of the sensors to the steel bridge beams. This sensor-mounting approach avoided issues relating to epoxies or mechanical clamping systems and worked well in terms of rapid deployment and recovery from the snooper truck. The approximate locations of the sensors are presented in Fig. 2. Each sensor collected 46 h of data on each axis. Meteorological data were also collected during the course of the experiment. Launches were constrained by the training schedule for Beale AFB with launches occurring at dawn and dusk on November 14 and 15 from the Sutter County Airport for a total of four launches. Analysis of the data allowed for identification of times most favorable for detection of signals from the bridge of interest.

Analysis and Results

Data analysis was approached as a two-step process focused on identifying infrasound signals from the Br 18–0009. Preliminary processing using MatSeis-InfraTool (Hart 2004) identified times of coherent signal, defined as signal present on at least three of five sensors in the array, across each array, with back azimuths corresponding to Br 18–0009. These results were used in conjunction with meteorological data obtained from the radiosonde launches to improve additional analysis. Following identification of the most promising times based on temperature inversions identified through analysis of data collected during radiosonde launches, manual processing of the data including filtering, Fourier analysis, and frequency-wave number (F-K) analysis was completed on the identified times utilizing Geotool, a software package (Coyne and Henson 1995). Initial data-analysis efforts yielded no detection of the low-energy signals at FRB during the initial times of interest (temperature inversions) identified in the collected radiosonde data, and the FRA array was located in close proximity to industrial sources that will require more sophisticated signal-processing techniques to eliminate. The data from March 1, 2014 at 0900–1000 Coordinated Universal Time (UTC) FRC array were the most promising; thus, the remainder of the analysis presented here was geared toward the processing of the FRC infrasound data from that time period array and the on-structure accelerometers from the second experiment.

Meteorological Considerations

Analysis of infrasound data requires an understanding of the prevailing meteorological conditions of the area at the time of the experiment, because these conditions will directly impact infrasound propagation. Considerations include temperature profile, wind speed and direction, and understanding of the layers of the atmosphere. The atmosphere is composed of four main layers: the troposphere (0–12 km), stratosphere (12–50 km), mesosphere (50–80 km), and thermosphere (80–320 km) (Evers and Haak 2009; Hedlin et al. 2012). A signal propagates from the source and spherically radiates upward through the atmosphere until a point is reached at which the effective sound speed, based on temperature, wind direction, and wind speed, is higher than at the origin of the signal. At this point, the signal turns and is refracted back down to the receiver. The phase velocity of the signal as it moves across the array is indicative of the atmospheric layer at which the signal turned. A higher phase velocity across the array indicates a higher turning point of the signal. Higher turning points are typically associated with long propagation paths.
The close proximity of the arrays to Br 18–0009 is considered local propagation, roughly defined as less than 50 km (McKenna et al. 2012); therefore, atmospheric effects are limited to the troposphere (<10 km). Analysis of the radiosonde data showed no significant winds, thereby reducing the sound speed [Eq. (1)] (McKenna et al. 2008; McKisic 1997) to adiabatic sound speed (without wind vector component)
Ceff= Ct+n·v
(1)
Ct20.07T
where T = absolute temperature (K); and n·v = component of wind speed in the direction of propagation.
Given the local propagation distance expected with the source-receiver spacings, it can be assumed that the wavefront propagated as a horizontal plane, with a phase velocity observed at the array approximately equal to the observed sound speed near the ground. The adiabatic sound speeds were calculated for data collected by each balloon launch with a maximum value of approximately 344 m/s. This value was later used in the frequency-wave number (F-K) analysis.

Initial Processing

InfraTool is an infrasound analysis tool within MatSeis and uses frequency-slowness processing to devolve data collected across an array into correlation, back azimuth, and slowness (Hart 2004). Two analysis schemes were devised to further analyze the output for determination of the presence of a signal, the inverse slope method and the Hough Transform, both of which were discussed by Hart (2004). InfraTool was originally designed to identify impulsive sources within uncorrelated noise but has proven to be a useful tool for preliminary processing to identify times of coherent continuous wave packetized signals and identify back azimuths. This allowed for rapid assessment of large quantities of data to determine times of interest for further processing.
For each of the arrays, InfraTool was used to process each hour of the 7-day data-collection period. InfraTool uses a Butterworth filter, with the number of poles and passbands established by the user. A Type I Chebyshev filter (investigated within Matlab) was considered for this analysis; however, too much ripple was introduced in the specified passband. The Butterworth filter is effectively a subset of the Chebyshev filter with zero allowed ripple. Butterworth filters were utilized here, because they give as flat a frequency response as possible in the passband of interest. In this experiment, a three-pole filter was specified to achieve fast roll-off while minimizing signal ringing. Each hour was processed twice, once with a passband from 0.5 to 10 Hz and again with a passband from 0.5 to 4 Hz. These bands were both well within the range defined as infrasound, with 0.5–10 Hz being a broader sweep and 0.5–4 Hz corresponding to a range comparable with frequencies seen for bridge modes in the literature. The window specified was a 10-s window with 50% overlap. Times of high correlation with a back azimuth in the direction of the bridge from each array for each 1-hour block were noted as possible detections and, thus, times of interest for further analysis. The March 1, 2014, 0900–1000 UTC time block from the FRC array contained high-correlation values (>0.8) with back azimuths consistently pointing at the source of interest and phase velocity approximately the measured sound speed.

Manual Processing

Following the identification of the time of interest (March 1, 2014, 0900–1000 UTC) for FRC through InfraTool and analysis of meteorological data, infrasound data were manually processed with Geotool, a software package for visualization and characterization of seismic and acoustic data (Coyne and Henson 1995). This processing involved filtering the data, handpicking coherent signals, completing a Fourier analysis to identify modes and confirm that the modes are in the infrasound passband, and an F-K analysis to confirm the back azimuth from which the signal originated (McKenna et al. 2009c, 2012; McComas et al. 2016).
First, the time series was filtered with a causal three-pole 1- to 10-Hz bandpass Butterworth filter based on a literature review that indicated the first several modes of long-span bridges are often under 5 Hz (Cunha et al. 2013; Conte et al. 2008; Brownjohn et al. 2010; Ren et al. 2005; Caicedo et al. 2001; Pietrzko et al. 1996; Hsieh et al. 2006; Morassi and Tonon 2008; Briaud et al. 2011; Foti and Sabia 2011; Jianxin et al. 2013; Chen et al. 2014; Ko et al. 2010; Lee et al. 2012) as well as a previous experiment using infrasound for bridge detection (McKenna et al. 2009a, c). The data were then manually processed to identify coherent data signals, such as those seen in the works by Donn et al. (1974), McKenna et al. (2009a, c), and McComas et al. (2016), which were defined as a signal present on at least three of the five infrasound sensors [Fig. 3(a)]. Examples of the bridge signal packets are highlighted with gray boxes.
Fig. 3. (Color) (a) Butterworth bandpass 1- to 10-Hz four-pole filtered infrasound with a delay and sum beam former applied to align coherent signals arriving at FRC, with bridge signal packets highlighted in gray; (b) Fourier analysis of the first four highlighted signal packets, 09:02:43–09:02:59 UTC, March 1, 2014, calculated with Welch’s Fourier analysis method using a 16-s (greater than 10 times longer than lowest expected frequency) Hanning window with 75% overlap; (c) Fourier analysis completed over a 1-h block (0900–1000, March 1, 2014, UTC) to show the persistent nature of bridge observations calculated with Welch’s Fourier analysis method using a 16-s Hanning window with 75% overlap; and (d) Fourier analysis of accelerometer data completed over a 1-h block with Welch’s Fourier analysis method using a 16-s Hanning window with 75% overlap.
When a coherent packet was identified, then the characteristics of the signal were noted [e.g., duration, frequency content, back azimuth to source, and signal-to-noise ratio (SNR)]. Fourier analysis was completed for each element of the FRC array to identify the coherent frequency in the signal packet window [Fig. 3(b)]. If coherent frequencies were observed, then F-K analysis was completed to determine back azimuth to the source and SNR. Visualization of the aforementioned signal processing scheme is presented in Fig. 3, which highlights the time series and frequency analysis. Fig. 3 was created in Matlab, because Geotool does not provide an adequate export capability for generating figures. The coherent continuous wave signal packets were highlighted in the time series data through the use of a delay and sum beamformer to align the time series in the direction of 222°, aligning with the main bridge span [Fig. 3(a)] (Rost and Thomas 2002).
The Fourier analysis for Fig. 3 was completed using Welch’s method, which is a nonparametric method used to estimate the power in a signal at different frequencies. Implementation is typically achieved by dividing a time domain signal into smaller overlapping and windowed segments that are converted to the frequency domain using the Fourier transform and then averaged at each frequency (Welch 1967). This approach has the benefit of typically reducing noise in the spectral estimate with some loss in frequency resolution due to the shortened time windows. This method was utilized to determine the coherent frequency content of a short (09:02:43–09:02:59 UTC, March 1, 2014) set of continuous wave packets and 1-h window (0900–1000 UTC, March 1, 2014), with both processed using 16-s windows and 75% overlap. The 16-s window was selected because it was 10 times longer than the lowest frequency of interest. The short-duration Fourier analysis identified the prominent frequency content of those signal packets [Fig. 3(b)], with the hour-long analysis demonstrating the persistent nature of the signal [Fig. 3(c)].
F-K analysis is an array-processing method by which the complete slowness vector, composed of both the back azimuth and the horizontal slowness, can be determined, with slowness being defined as the inverse of the apparent velocity for the wave front of the signal as it moves across the array, and the back azimuth being defined as the angle of the wave front arriving at the array measured between north and the direction to the source in degrees (Rost and Thomas 2002). This analysis is only applicable for short time windows, because longer time frames may contain several different phases, making differentiation and localization of the back azimuth difficult to differentiate between phases (Rost and Thomas 2002). The packetized signals that were investigated here occurred in short bursts, so F-K analysis was applicable in this case. Within Geotool, users are allowed to set the maximum slowness to be considered. This parameter was set based on the observed adiabatic sound speed. It is important to note that adiabatic sound speed and phase velocity are two different parameters, but for suspected direct arrivals where the source-to-receiver distance is in the local propagation range, it is acceptable to assume that the two values are approximately equal.
Conversion of effective sound speed (in meters per second) to slowness (in seconds per degree) yielded a value of 310 s/degree. Inputting this parameter as the maximum slowness for the F-K analysis and hand tuning the results for an apparent velocity of 0.36 km/s, corresponding to the 361 m/s determined previously, yielded the results presented in Fig. 4.
Fig. 4. (Color) F-K analysis for a 6.5-s duration starting at 09:02:43 and hand tuned to an apparent velocity of 0.36 km/s.
The contours of the plot indicate relative power, with the back azimuth being associated with the highest relative power on the plot that matches the slowness for the expected return of the signal as it is directed back to the earth. The results indicated a back azimuth of 222° from FRC, which corresponded to Br 18–0009. Fig. 5 presents the range of back azimuths from the FRC array to Br 18–0009, which ranged from 209 to 223°. The back azimuth of 222° corresponded to one of the longest spans of the bridge, which was expected because the long spans were able to displace more air, generating the infrasonic waves we detected. The result of the F-K analysis and the coherent frequencies observed indicated the hypothesized source of Br 18–0009.
Fig. 5. Plotted range of back azimuths from FRC that aligned with Br 18–0009.

Accelerometer and Near-Field Infrasound Sensor Analysis and Results

Reviewing the data comprehensively, it appeared that the periods of increased traffic, such as rush hour, transferred more energy into the higher harmonics, whereas the less busy periods were dominated more by the natural mode, near 2 Hz. This is thought to be a product of the traffic forcing function and the increased rate of the sources (i.e., the frequency of occurrence of traffic wheel loads). Analysis of a 46-h data-collection window, from approximately 1100 Pacific Standard Time (PST) on November 16 to 0930 PST on November 18, 2015, indicated that traffic was not as efficient at exciting the lowest modes of the structure.
A number of different approaches were pursued to visualize the bridge’s response as a function of traffic characteristics. When vehicles crossed the expansion joint, the bridge in the area of the accelerometer was excited. These were counted as vehicle detections. Acceleration levels related to these detections, which are generally proportional to the wheel load impacts, were binned based on several peak-amplitude levels. Fourier analyses using Welch’s method with a 6-s window and zero overlap were then completed for each detection, and their spectral responses were averaged to get an estimated response as a function of detected peak-acceleration levels. An example composite spectrum showing this change in energy as a function of traffic peak-acceleration levels is presented in Fig. 6. The purple trace is the averaged composite for traffic peak impacts exceeding 0.075 G, and the red trace is averaged impacts under 0.075 G. The blue trace is from averaged impacts over 0.04 G, and the yellow trace is due to averaged impacts below 0.04 G. From this visualization, it is clear that some traffic loads produced up to 50 dB more response than others.
Fig. 6. (Color) Response of Sensor 4 accelerometer as a function of increased wheel load impact.
Fig. 3(d) presents Welch’s power spectral density estimate for a representative accelerometer deployed on Bridge span 22 corresponding to 0900–1000 UTC on November 17, 2015. The parameters for Welch’s method aligned with the infrasound processing, 16-s windows with 75% overlap. To minimize errors due to variations in traffic type, location, and rates of excitation, long observations over the 46-h deployment were compared with shorter 1-h durations using Welch’s spectral estimation method.
For both spans, the strongest vibrations were observed at the sensors located closest to the bridge deck’s outer edge. This was expected because that location was only constrained on one side and could have more motion. For Span 21, the dominant low frequency was 2.27 Hz, whereas the shorter Span 22 had a dominant response at approximately 2.32 Hz. For comparison purposes, an infrasound sensor was also located on the ground at the bridge. Welch’s power spectral estimate was performed using the same parameters as described previously for the infrasound analysis; however, there were some potential issues with the data from this sensor. To limit spatial aliasing in processing, infrasound sensors should be deployed at a distance to allow for at least one wavelength to be completed before the signal reaches the sensor. Because of this, the lowest modes of the sensor at the bridge could not be properly resolved. Infrasound technology is actually more efficient at detection of the lowest modes when it is used at a distance from the structure of interest. In addition, the close proximity of the sensor at the bridge means there was a possibility that some of the frequencies recorded were due to seismic coupling.

Comparison of Infrasound and Accelerometer Results

In comparing the results obtained from the analysis of the accelerometers and the infrasound data at array FRC located 2.6 km from the bridge, presented in Fig. 3, there was good agreement in the frequencies, seen in the highlighted bands in Figs. 3(b–d). Comparison of the accelerometer data [Fig. 3(d)] and the data from the FRC array located 2.6 km from Br 18–0009 [Figs. 3(b and c)] indicates that the modes at approximately 0.9 and 1.6 Hz were faint in the accelerometer data. One hypothesis is that the higher modes excited by the traffic may have overwhelmed the lower modes on the accelerometers, as in Conte et al. (2008). Conte et al. (2008) performed a series of experiments on the Alfred Zampa Memorial Bridge in California prior to the bridge opening to traffic. The bridge was instrumented with a series of accelerometers, and a series of tests were completed including forced vibration tests, consisting mostly of traffic loads, and ambient vibration tests, consisting mostly of excitation due to wind. Test results indicated that the forced vibration tests were better able to excite the higher-frequency modes, making the lower modes more difficult to identify. The ambient vibration tests, utilizing mostly wind, were shown to be more informative for identification of the lowest modes (Conte et al. 2008).
This observation holds for the current data as well. The on-structure accelerometers were very good at picking up the local effects of rapid transient loadings (wheel loads) at their points of placement. Although these effects overwhelmed the lower-order modes at a given point, they did not contribute as much to the gross structural motion that moved large masses of air, creating infrasound signals. The infrasound data presented in Fig. 3 corresponded to approximately 0900 UTC. In local time, that corresponds to 0200 PST, which would be a time with less traffic. A potential benefit of infrasound for structural health monitoring is that it naturally filters out the higher-order transient signals while emphasizing lower-order primary vibration modes that are most closely tied to global structural changes.

Implications

Analyses and comparisons of the presented infrasound and accelerometer data indicate the feasibility of using infrasound for noncontact, nonline-of-sight monitoring of bridges. The remainder of this section discusses the results and their implications for the technology, including monitoring for global structural change and postevent prioritization for inspection.
Initial research investigating the feasibility of infrasound for remote bridge monitoring indicated that some global characteristics, such as scour, could be more readily detected with the use of infrasound (Whitlow et al. 2012, 2013). Scour, as defined by the Federal Highway Administration, is the “result of the erosive action of flowing water, excavating and carrying away material from the bed and banks of streams and from around the piers and abutments of bridges,” and is the most common cause of bridge failure in the United States (Hunt 2009; Richardson and Davis 2001). According to the 2017 NBI data, there are currently 47,971 bridges classified as scour critical or with unknown foundations (FHWA 2017). The NBI standard defines a scour-critical bridge as one with a foundation element that has been determined to be unstable for the observed or evaluated scour condition (FHWA 2004). Bridges classified as having an unknown foundation have not been evaluated for scour because of a lack of information to complete the scour analysis. In both cases, there is a significant need for persistent monitoring of the bridge.
Ko et al. (2010) presented a paper in 2010 at the International Conference on Scour and Erosion detailing the use of vibration measurement, and more specifically frequency change, to evaluate scour at several bridges in Taiwan. Utilizing both traffic-induced and ambient vibration data on multiple bridges with known scour problems in conjunction with a finite-element model, the group was able to show a clear reduction in the natural frequency of each structure due to a decrease in the stiffness of the structure as the scour increased (Ko et al. 2010). These results were further validated by Lee et al. (2012) in the testing of an advanced monitoring system to monitor scour at critical bridges.
A case study conducted by Foti and Sabia (2011) investigated the use of frequency change to monitor ongoing scour at a bridge in Turin, Italy. Testing was completed for the bridge both before and after a retrofit of the affected pier. Results concluded that scour detection through frequency change was possible, and determination of depth of scour might also be a possibility. Prendergast et al. (2013) furthered this concept when they conducted laboratory and field testing on a pile exposed to scour to investigate the change in dynamic response with progressive scour. Whereas the study presented a simple case of a single pile, the results showed that scour detection and assessment of scour depth are possible using a change in the natural frequency as a result of stiffness decreasing. This method was further expanded to a full bridge subjected to traffic loading in a study by Prendergast et al. (2016).
The studies by Ko et al. (2010), Lee et al. (2012), and Foti and Sabia (2011) and the method proposed by Prendergast et al. (2016) all utilized or proposed to utilize on-structure instrumentation, leaving them vulnerable to the same issues all other types of on-structure or near-structure instrumentation are subject to, such as damage from high-velocity flows, debris, ice forces, sediment loading, severe water temperatures, and vandalism (Hunt 2009). Because infrasonic monitoring is noncontact, nonline of sight, and still allows for detection of the natural frequencies of the structure, the aforementioned issues can be avoided while still providing a method for monitoring of global structural changes, such as scour.
This technology also has implications for prioritization of structures for inspection and use postevent. In the hours and days immediately following an extreme event, whether natural or artificial, that could cause damage to infrastructure, understanding which infrastructure is still sound is vital for first responders arriving for rescue operations and later for moving supplies into affected areas. Infrasonic bridge monitoring could provide a means to quickly establish prioritization for inspection postevent to best utilize limited resources. Fig. 7 provides examples of bridge scour taken after storm damage from hurricanes Harvey in Texas and Irma in Florida in 2017. The ability to detect damage of this nature via infrasound could have some potential value for disaster recovery purposes. These photos were taken during recovery efforts. Hurricane Irma photos were taken during visual go or no-go inspections for heavy equipment and rescue team entry into damaged areas.
Fig. 7. (Color) Bridges damaged by scour (or wave action) during hurricanes: (a) bridges after Harvey in 2017 (images courtesy of Janice M. Maaskant); and (b) bridges after Irma in 2017 (images courtesy of Rob Garner).
The strength of this technology lies in the ability to persistently and remotely monitor a structure without line of sight to that structure. The usefulness of this technology as a supplementary bridge monitoring technique would be greatly reduced if it required either a finite-element model for prediction of bridge frequencies or on-structure validation for the same. To establish a baseline by which to compare frequency changes, it is anticipated that bridge inspections in conjunction with initial monitoring could be used to determine the baseline for a bridge. Because all bridges are required to be inspected every 2 years by law (FHWA 2004), and more often in some cases with identified deficiencies, we can gain a good understanding of the initial condition of a structure when monitoring begins. Changes beyond the frequencies detected at that point would be the indicators that additional inspection is warranted.
As a postevent prioritization tool, a network of arrays deployed in an area to monitor a series of bridges deemed critical to transportation or in need of more frequent monitoring from some identified issue can establish a baseline for the bridges being monitored. A comparison of postevent bridge frequencies with the baseline, pre-event frequencies could potentially allow for bridges to be differentiated into groups based on whether a change in frequencies is detected or if the frequencies have stayed the same.

Summary and Conclusions

The purpose of the experiments documented in this paper was to validate the use of infrasound as a means of remote, noncontact sensing for bridge monitoring. Two experiments were completed on Br 18–0009, a steel two-girder bridge located in northern California. The first experiment collected infrasound data using array infrasound sensors. The second experiment sought to validate the frequencies determined through analysis of infrasound data by placing accelerometers on the main spans of bridge, although in future validation studies, high-precision accelerometers will be used to reduce potential errors in time stamping and digitization. The accelerometer placement was chosen to capture the lowest-mode structural vibrations because the movement of those spans would be the most likely candidate for generation of infrasonic waves, because they typically represent the vertical modes of the bridge, allowing the bridge deck to become engaged. Comparison of the frequency data from analysis of the infrasound data and the accelerometer data showed excellent agreement, validating the use of infrasonic monitoring as a new method of noncontact, persistent bridge monitoring. This technology could serve as a complement to existing monitoring practices while also having implications for postevent prioritization of structures.

Acknowledgments

The valuable assistance of Kevin Flora and Caltrans for collaboration concerning the Br 18–0009 is gratefully acknowledged. Special thanks are extended to Beale AFB, Marysville City Hall, and Sutter County Airport for allowing instrumentation deployment. Thanks are extended to Janice M. Maaskant and Rob Garner for the use of their Hurricane Harvey and Hurricane Irma photos, respectively. Thanks are also extended to Rodney Gonzalez Rivera, Theodore A. Lee III, and Henry Diaz Alvarez for assistance in figure generation. Permission to publish was granted by the Director, Geotechnical and Structures Laboratory, US Army Engineer Research and Development Center. This work was funded by the Assistant Secretary of the Army (Acquisition, Logistics, and Technology) [ASA(ALT)] with portions funded under 62784/T40/24, 62784/T40/24V, and 62784/T40/46.

References

Bass, H. E., J. Bhattacharyya, M. A. Garces, M. Hedlin, J. V. Olson, and R. L. Woodward. 2006. “Infrasound.” Acoust. Today 2 (1): 9–19. https://doi.org/10.1121/1.2961130.
Bedard, A., and T. Georges. 2000. “Atmospheric infrasound.” Acoust. Aust. 28 (2): 47–52.
Briaud, J.-L., S. Hurlebaus, K. A. Chang, C. Yao, H. Sharma, O.-Y. Yu, C. Darby, B. E. Hunt, and G. R. Price. 2011. Realtime monitoring of bridge scour using remote monitoring technology. FHWA/TX-11/0-6060-1. Austin, TX: Texas Dept. of Transportation Research and Technology Implementation Office.
Brownjohn, J. M. W., F. Magalhaes, E. Caetano, and A. Cunha. 2010. “Ambient vibration re-testing and operational modal analysis of the Humber Bridge.” Eng. Struct. 32 (8): 2003–2018. https://doi.org/10.1016/j.engstruct.2010.02.034.
Caicedo, J. M., J. Marulanda, P. Thomson, and S. J. Dyke. 2001. “Monitoring of bridges to detect changes in structural health.” In Vol. 1 of Proc., 2001 American Control Conf., 453–458. New York: IEEE.
Chen, S.-E., W. Liu, K. Dai, H. Bian, and E. Hauser. 2011. “Remote sensing for bridge monitoring.” In Proc., 2011 Condition, Reliability, and Resilience Assessment of Tunnels and Bridges, Geotechnical Special Publication 214, 118–125. Reston, VA: ASCE.
Chen, C.-C., W.-H. Wu, F. Shih, and S.-W. Wang. 2014. “Scour evaluation for foundation of a cable-stayed bridge based on ambient vibration measurements of superstructure.” NDT and E Int. 66: 16–27. https://doi.org/10.1016/j.ndteint.2014.04.005.
Chopra, A. K. 2012. Dynamics of structures: Theory and applications to earthquake engineering. 4th ed. Upper Saddle River, NJ: Prentice Hall.
Christie, D. R., and P. Campus. 2010. “The IMS infrasound network: Design and establishment of infrasound stations.” In Infrasound monitoring for atmospheric studies, edited by A. Le Pichon, E. Blanc, and A. Hauchecorne, 29–75. Dordrecht, Netherlands: Springer.
Conte, J. P., X. He, B. Moaveni, S. F. Masri, J. P. Caffrey, M. Wahbeh, F. Tasbihgoo, D. H. Whang, and A. Elgamal. 2008. “Dynamic testing of Alfred Zampa Memorial Bridge.” J. Struct. Eng. 134 (6): 1006–1015. https://doi.org/10.1061/(ASCE)0733-9445(2008)134:6(1006).
Coyne, J. M., and I. Henson. 1995. Geotool sourcebook: User’s manual. Huntsville, AL: Teledyne Brown Engineering.
Cunha, A., E. Caetano, F. Magalhães, and C. Moutinho. 2013. “Recent perspectives in dynamic testing and monitoring of bridges.” Struct. Control Health Monit. 20 (6): 853–877. https://doi.org/10.1002/stc.1516.
Diaz-Alvarez, H., M. H. McKenna, and P. Mlakar. 2009. Infrasound assessment of infrastructure. Report 1: Field testing and finite element analysis for railroad bridge A.B 0.3. ERDC/GSL TR-09-16. Vicksburg, MS: US Army Engineer Research and Development Center.
Donn, W. L., N. K. Balachandran, and G. Kaschak. 1974. “Atmospheric infrasound radiated by bridges.” J. Acoust. Soc. Am. 56 (5): 1367–1370. https://doi.org/10.1121/1.1903451.
Evans, J. R., R. Allen, A. Allen, E. Chung, R. Cochran, M. Hellweg, and J. Lawrence. 2014. “Performance of several low cost accelerometers.” Seismol. Res. Lett. 85 (1): 147–158. https://doi.org/10.1785/0220130091.
Evers, L. 2005. “Infrasound monitoring in the Netherlands.” J. Acoust. Soc. Neth. 176: 1–11.
Evers, L. G., and H. W. Haak. 2009. “The characteristics of infrasound, its propagation and some early history.” In Infrasound monitoring for atmospheric studies, edited by A. Le Pichon, E. Blanc, and A. Hauchecorne, 3–27. Dordrecht, Netherlands: Springer.
Farrar, C. R., T. Duffey, P. J., Cornwell, and S. W. Doebling. 1999. “Excitation methods for bridge structures.” In Vol. 1 of Proc., Society for Experimental Mechanics, Inc., 17th Int. Modal Analysis Conf., 1063–1068. Bethel, CT: Society for Experimental Mechanics.
FHWA (Federal Highway Administration). 2004. “National bridge inspection standards (NBIS).” Accessed December 7, 2017. http://www.fhwa.dot.gov/bridge/nbis.htm.
FHWA (Federal Highway Administration). 2017. “National Bridge Inventory (NBI).” Accessed February 11, 2018. https://www.fhwa.dot.gov/bridge/nbi/ascii.cfm.
Foti, S., and D. Sabia. 2011. “Influence of foundation scour on the dynamic response of an existing bridge.” J. Bridge Eng. 16 (2): 295–304. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000146.
Fukuda, Y., M. Q. Feng, Y. Narita, S. Kaneko, and T. Tanaka. 2013. “Vision-based displacement sensor for monitoring dynamic response using robust object search algorithm.” IEEE Sens. J. 13 (12): 4725–4732. https://doi.org/10.1109/JSEN.2013.2273309.
Gentile, C., and A. Cabboi. 2015. “Vibration-based structural health monitoring of stay cables by microwave remote sensing.” Smart Struct. Syst. 16 (2): 263–280. https://doi.org/10.12989/sss.2015.16.2.263.
Harris, D. K., C. N. Brooks, and T. M. Ahlborn. 2016. “Synthesis of field performance of remote sensing strategies for condition assessment of in-service bridges in Michigan.” J. Perform. Constr. Facil. 30 (5): 04016027. https://doi.org/10.1061/(ASCE)CF.1943-5509.0000844.
Hart, D. 2004. Automated infrasound signal detection algorithms implemented in MatSeis-Infra tool. Washington, DC: US Dept. of Energy.
Hedlin, M. A. H., K. Walker, D. P. Drob, and C. D. de Groot-Hedlin. 2012. “Infrasound: Connecting the solid earth, oceans, and atmosphere.” Ann. Rev. Earth Planet. Sci. 40 (1): 327–354. https://doi.org/10.1146/annurev-earth-042711-105508.
Hsieh, K. H., M. W. Halling, and P. J. Barr. 2006. “Overview of vibrational structural health monitoring with representative case studies.” J. Bridge Eng. 11 (6): 707–715. https://doi.org/10.1061/(ASCE)1084-0702(2006)11:6(707).
Hunt, B. E. 2009. NCHRP synthesis 396: Monitoring scour critical bridges. Washington, DC: Transportation Research Board.
Jianxin, L., W. Ying, Z. Meichun, Z. Zhihong, Z. Xinhua, G. Jianhong, L. Pei, and T. Wen. 2013. “Static and dynamic behavior of the high-pier and long-span continuous rigid frame bridge.” Adv. Mater. Res. 639–640: 474–480.
Ko, Y. Y., W. F. Lee, W. K., Chang, H. T. Mei, and C. H. Chen. 2010. “Scour evaluation of bridge foundations using vibration measurement.” In Proc., Int. Conf. on Scour and Erosion 2010 (ISCE-5). Reston, VA: ASCE. https://doi.org/10.1061/41147(392)88.
Lee, W. F., T. T. Cheng, C. K., Huang, C. H., Chen, and J. Y. C. Lai. 2012. “Development and verification of an advanced monitoring system for bridge foundations under scour.” In Proc., Int. Conf. on Scour and Erosion 2012 (ICSE-6), 445–452. London: Int. Society for Soil Mechanics and Geotechnical Engineering.
McComas, S. L., M. E. Pace, H. Diaz Alvarez, C. Simpson, and M. H. McKenna. 2016. Persistent monitoring of urban infrasound phenomenology. Report 2: Investigation of structural infrasound signals in an urban environment. ERDC/GSL TR-15-5. Vicksburg, MS: US Army Engineer Research and Development Center.
McKenna, M. H., R. G. Gibson, B. E. Walker, J. McKenna, N. W. Winslow, and A. S. Kofford. 2012. “Topographic effects on infrasound propagation.” J. Acoust. Soc. Am. 131 (1): 35–46. https://doi.org/10.1121/1.3664099.
McKenna, M. H., A. F. Lester, and S. McComas. 2009a. Infrasound assessment of infrastructure. Report 2: Experimental infrasound measurements of rail road bridge A.B 0.3. ERDC/GSL TR-09-16. Vicksburg, MS: US Army Engineer Research and Development Center.
McKenna, M. H., S. McComas, A. Lester, and P. Mlakar. 2009b. “Infrasound measurement of a railroad bridge.” In Proc., 22nd Annual Symp. on the Application of Geophysics to Engineering and Environmental Problems (SAGEEP), 669–678. Denver, CO: Environmental and Engineering Geophysical Society.
McKenna, M. H., B. W. Stump, and C. Hayward. 2008. “Effect of time-varying tropospheric models on near-regional and regional infrasound propagation as constrained by observational data.” J. Geophys. Res. 113: D11111. https://doi.org/10.1029/2007JD009130.
McKenna, M. H., S. Yushanov, K. Koppenhoefer, and J. R. McKenna. 2009c. Infrasound assessment of infrastructure. Report 3: Numerical simulation of structural acoustic coupling and infrasonic propagation modeling for rail road bridge A.B 0.3. ERDC/GSL TR-09-16. Vicksburg, MS: US Army Engineer Research and Development Center.
McKisic, J. M. 1997. Infrasound and the infrasonic monitoring of atmospheric nuclear explosions: A literature review. PL-TR-97-2123. Bedford, MA: Phillips Laboratory, Directorate of Geophysics, Hanscom Air Force Base.
Morassi, A., and S. Tonon. 2008. “Dynamic testing for structural identification of a bridge.” J. Bridge Eng. 13 (6): 573–585. https://doi.org/10.1061/(ASCE)1084-0702(2008)13:6(573).
Pietrzko, S. J., R. Cantieni, and Y. Deger. 1996. “Modal testing of a steel/concrete composite bridge with a servo-hydraulic shaker.” In Proc., 14th Int. Modal Analysis Conf., 91–98. Bethel, CT: Society for Experimental Mechanics.
Prendergast, L. J., K. Gavin, and C. Reale. 2016. “Sensitivity studies on scour detection using vibration-based systems.” Transp. Res. Procedia 14: 3982–3989. https://doi.org/10.1016/j.trpro.2016.05.495.
Prendergast, L. J., D. Hester, K. Gavin, and J. J. O’Sullivan. 2013. “An investigation of the changes in the natural frequency of a pile affected by scour.” J. Sound Vib. 332 (25): 6685–6702. https://doi.org/10.1016/j.jsv.2013.08.020.
Ren, W. X., X. L. Peng, and Y. Q. Lin. 2005. “Experimental and analytical studies on dynamic characteristics of a large span cable-stayed bridge.” Eng. Struct. 27 (4): 535–548. https://doi.org/10.1016/j.engstruct.2004.11.013.
Richardson, E. V., and S. R. Davis. 2001. Evaluating scour at bridges: Hydraulic engineering circular No. 18. 4th ed. FHWA NHI 01-001. Washington, DC: Federal Highway Administration.
Rost, S., and C. Thomas. 2002. “Array seismology: Methods and applications.” Rev. Geophys. 40 (3): 1–2.
Salawu, O. S. 1997. “Detection of structural damage through changes in frequency: A review.” Eng. Struct. 19 (9): 718–723. https://doi.org/10.1016/S0141-0296(96)00149-6.
Simpson, C. P., S. L. McComas, R. D. Whitlow, K. M. McLaughlin, M. H. McKenna, and B. G. Quinn. Forthcoming. SIAM deployment and documentation: A guide and case study. ERDC/GSL TR. Vicksburg, MS: US Army Engineer Research and Development Center.
Sorrells, G. G., E. T. Herrin, and J. L. Bonner. 1997. “Construction of regional ground truth databases using seismic and intrasound data.” Seismol. Res. Lett. 68 (5): 743–752. https://doi.org/10.1785/gssrl.68.5.743.
Vaghefi, K., R. C. Oats, D. K. Harris, T. M. Ahlborn, C. N. Brooks, K. A. Endsley, C. Roussi, R. Shuchman, J. W. Burns, and R. Dobson. 2012. “Evaluation of commercially available remote sensors for highway bridge condition assessment.” J. Bridge Eng. 17 (6): 886–895. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000303.
Welch, P. 1967. “The use of fast Fourier transform for the estimation of power spectra: A method based on time averaging over short, modified periodograms.” IEEE Trans. Audio Electroacoust. 15 (2): 70–73. https://doi.org/10.1109/TAU.1967.1161901.
Whitlow, R. D., M. H. McKenna, O. D. S. Taylor, R. D. Costley, M. C. L. Quinn, and H. Diaz-Alvarez. 2012. Infrasound assessment of infrastructure. Report 4: Investigation of scour using infrasound for railroad bridge A.B. 0.3. ERDC/GSL TR-09-16. Vicksburg, MS: US Army Engineer Research and Development Center.
Whitlow, R. D., O. D. S. Taylor, M. H. McKenna, and M. C. L. Quinn. 2013. Infrasound assessment of infrastructure. Report 6: Scour detection and riverine health assessment using infrasound. ERDC/GSL TR-09-16. Vicksburg, MS: US Army Engineer Research and Development Center.

Information & Authors

Information

Published In

Go to Journal of Bridge Engineering
Journal of Bridge Engineering
Volume 24Issue 5May 2019

History

Received: Apr 2, 2018
Accepted: Oct 1, 2018
Published online: Feb 18, 2019
Published in print: May 1, 2019
Discussion open until: Jul 18, 2019

Authors

Affiliations

R. Danielle Whitlow [email protected]
P.E.
Research Civil Engineer, US Army Engineer Research and Development Center; Doctoral Candidate, Mississippi State Univ., 3909 Halls Ferry Rd., Vicksburg, MS 39180 (corresponding author). Email: [email protected]
Richard Haskins
Research Electronics Engineer, US Army Engineer Research and Development Center, 3909 Halls Ferry Rd., Vicksburg, MS 39180.
Sarah L. McComas
Research General Engineer, US Army Engineer Research and Development Center, 3909 Halls Ferry Rd. Vicksburg, MS 39180.
C. Kennan Crane, Ph.D., M.ASCE
P.E.
Research Civil Engineer, US Army Engineer Research and Development Center, 3909 Halls Ferry Rd., Vicksburg, MS 39180.
Isaac L. Howard, Ph.D., F.ASCE
P.E.
Materials and Construction Industries Chair, Dept. of Civil and Environmental Engineering, Mississippi State Univ., 501 Hardy Rd., PO Box 9546, Mississippi State, MS 39762.
Mihan H. McKenna, Ph.D.
Senior Scientific Technical Manager, US Army Engineer Research and Development Center, 3909 Halls Ferry Rd., Vicksburg, MS 39180.

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share