In this paper, the results from using a large active ring laser interferometer as an infrasound detector are presented. On April 27, 2014, an EF4 tornado struck Central Arkansas and passed within 21 km of the ring laser interferometer. The tornado resulted in 16 fatalities and millions of dollars in damage. Using the ring laser to study the tornado infrasound produced results that qualitatively agree with several findings from a long-term study of weather generated infrasound by the National Oceanic and Atmospheric Administration. A Fast Fourier Transform of the ring laser output revealed a coherent frequency of approximately 0.94 Hz that lasted during the life of the storm. The 0.94 Hz frequency was initially observed 30 min before the funnel was reported on the ground. Infrasound signatures from four separate tornadoes are presented. In each case, coherent infrasound was detected at least 30 min before the tornado was reported on the ground. Examples of the detection of distant coherent acoustic-gravity waves from volcanoes and typhoons are also presented. In addition, buoyancy waves were recorded.

After Macek et al.1 demonstrated that a ring laser could detect rotation in the early1960s, the aerospace industry began developing and deploying small navigational ring lasers2–4 as replacements for mechanical gyroscopes. More recently, large ring laser interferometers have been developed to explore numerous geophysical effects. A good review of the history, design, and use of large active ring lasers is contained in an invited review article by Schreiber and Wells.4 In the current paper, the results from using large active ring laser interferometers to detect infrasound are presented. Acoustic vibrations below the lower limit of human hearing of approximately 20 Hz are typically labeled as infrasound. An overview of atmospheric infrasound is contained in a paper5 by Bedard and Georges. In the 1970s, National Oceanic and Atmospheric Administration (NOAA) began a long term study to determine if infrasound could help to improve warnings of severe weather events. In a 2005 paper,6 Bedard examined the results from this and other studies that were devoted to measuring acoustic energy generated by severe weather and related tornadoes. He concluded the results were consistent with a model by Abdullah7 that suggested the radial vibration modes in tornado vortices were a source of infrasound and that the infrasound frequencies were inversely proportional to the vortex diameter. Bedard6 cited Doppler radar measurements of an evolving tornado; the maximum values of the circulation were observed to descend over a period of ∼30 min. A tornado funnel was sighted when the maximum circulation was near the surface of the Earth. During the 22nd American Meteorological Society Conference on Severe Local Storms in October of 2004, two sessions8,9 were devoted to an infrasound network that was deployed in 2003. Several conclusions from this conference were based on the observations of numerous tornadoes: infrasound was usually observed well before the tornado funnel was reported on the ground; infrasound could travel long distances; vortex related signals usually last for 10 min or more; signals of 1 min or less were unlikely to be related to coherent vortex oscillations; and wind noise abatement systems improved tornado detection. During the first session,8 several ways a tornado could produce sound were discussed. Some of the possibilities include shear instabilities, boundary layer sound, fluid instabilities, and radial modes of vibration of the vortex core. Radial modes of vibrations of the vortex core were suggested as the most likely source. The focus6 of the NOAA studies was from 0.5 to 5 Hz. Although a ring laser interferometer is a totally different detection system, the results presented in this paper are in general agreement with the NOAA findings.

In an active ring laser interferometer, the excitation source is inside the cavity where it excites two counter-propagating waves. When the ring laser cavity is rotating, the two 632.8 nm He-Ne counter-propagating waves circulating in the cavity travel unequal path lengths.2–4 Unequal path lengths produce frequency differences between the two counter-propagating waves that result in a beat note.

The square ring laser is 5.65 m on a side. The cavity is excited by a 75 cm plasma tube that is sealed on each end with a Brewster window. Ambient pressure dry nitrogen created by evaporating liquid nitrogen from a well insulated 20 l Dewar fills the remainder of the cavity. Outgassing from the unbaked aluminum beam lines is slowly flushed into an oil bath. The oil bath prevents atmospheric contamination of the cavity. However, it does not isolate the cavity from atmospheric pressure variations and infrasound. The flush is essentially constant; it takes up to a month to evaporate the liquid in the Dewar. Consequently, the effect of any bias introduced by the flush is considered to be negligible. To help isolate the cavity from temperature expansions and contractions of the beam lines, soft air tight couplings are used to connect the beam lines to the mirror boxes and the plasma tube container. The soft couplings are nearly transparent to atmospheric pressure variations. The laser cavity's stability is partially dictated by the basement slab on which the laser is rigidly attached. The Sagnac frequency or beat note induced by Earth's rotation is given by the following4 relation:

(1)

In this relation, Δf is the Sagnac frequency and A is the area of the cavity. The symbol p represents the perimeter of the cavity, λ is the laser's wavelength, Ω is the rotation rate of the Earth, and θ is the angle between the normal to the laser cavity and the Earth's axis of rotation. The relation 4A/λp is commonly referred to as the scale factor. Since the colatitude of the ring laser is ∼55° north and each side is 5.69 m, Earth's rotation produces a beat note of ∼375 Hz. The beat note serves as a carrier frequency. When the ring laser cavity is perturbed, the carrier frequency (beat note) is frequency modulated (FM). Numerous geophysical phenomena can perturb a ring laser cavity and modulate the carrier frequency. For example, phenomena such as the Chandler wobble, diurnal polar motion, tidal tilts, and solid earth tides produce variations4 in θ. Rotational ground motion, introduced by seismic shear waves that add to or subtract from Ω, has been detected by several research teams.4,10–13 Variations in the cavity geometry can modulate the beat frequency.4 Also, infrasound can frequency modulate14,15 the carrier frequency. A phase locked loop16 is often used as a FM demodulation technique to recover the modulation (perturbing) signals. If the perturbing signals are sufficiently coherent, Fast Fourier Transforms (FFTs) can reveal their frequency spectra. The results reported in this paper were obtained using an alternate FM demodulation scheme. An experimental comparison between the phase locked loops and the alternate scheme produced equivalent results.14 

The demodulation system was built around a National Instruments Virtual Instrument (VI), “Extract Single Tone Information.” The VI takes a signal lasting for a certain interval of time and selects the single frequency in that interval with the highest amplitude. In this study, the digitized signal of 4000 samples/s was broken into 0.1 s intervals. The VI calculated the frequency with the highest amplitude in each 0.1 s interval. These results are initially stored in 10 min blocks. The stored data can then be used to show the variations of the carrier frequency with time. Typically, the ten minute data blocks are joined together creating extended data blocks. Each FFT measurement needs to include enough data to significantly exceed the Nyquist criteria. An FFT routine can then show the coherent modulation frequency spectrum in each data block.

Originally, the Arkansas ring laser was designed to be a relatively inexpensive seismic wave detector that did not require expensive turbomolecular vacuum pumps. To achieve this goal, a plasma tube and industrial grade dielectric mirrors were used. In Fig. 1, the dielectric mirror and its holder are shown in the mirror box. Figure 2 shows how the plasma tube is mounted. In addition, outgassing was flushed from the cavity using dry nitrogen. Initially, it was deployed to study shallow earthquakes associated with the underground disposal of waste water from hydraulic fracking operations in North Central Arkansas. When deployed, the interferometer not only detected regional earthquakes but also infrasound from hurricanes.

To verify that the ring laser was actually sensitive to infrasound, a controlled experiment using a portable Helmholtz acoustic resonator was set up in the basement containing the ring laser. As discussed by Kinsler,17 if the acoustic wavelength is greater than the dimensions of a Helmholtz acoustic resonator, the device behaves like a harmonic oscillator in one dimension. In this case, the Helmholtz resonator consists of a 55-gallon drum with a half-inch thick aluminum top that is connected to an aluminum duct 1-m long and 10.2 cm in diameter; the duct is open to the external environment. Air in the cavity (drum) acts as a spring, while the air in the duct acts as a mass; for a given geometry, a well-defined resonance frequency is created. An audio woofer between the volume and duct drives the combined system at resonance. The radiated infrasound from the duct opening was very small but sufficient for these tests. The resonator was located ∼3 m from the enclosed plasma tube. Care was taken to isolate the Helmholtz resonator from the floor by suspending it from the ceiling rafters so that the vibrations initially traveled through the air. At resonance, the Helmholtz acoustic generator emitted a very sharp frequency14 of 12.5 Hz. An FFT of the ring laser response14 agreed with the 12.5 Hz frequency generated by the Helmholtz acoustic resonator. Using FFTs, the coherent acoustic-gravity waves from hurricanes and explosive volcanic eruptions were subsequently detected. This paper reports the detection of tornadic infrasound as well as the acoustic-gravity waves from typhoons in the Pacific Ocean. FFTs are required to extract the comparative low level coherent acoustic responses from the more random higher intensity environmental noise.

FIG. 1.

Cavity mirror in the mirror box without its cover. The mirrors outside the mirror box form the combining optics. The reflectivity of the cavity mirror is greater than 99.7%. Small amounts of light from the counter-propagating laser beams are transmitted through the dielectric cavity mirror and directed by the mirrors outside the viewports onto the beam splitter. The vertical adjustor is part of the beam splitter. When the transmitted light is collimated and focused by the bottom mirror on the photodiode at the extreme left of the breadboard, a beat frequency proportional to the rotation rate of the cavity is created. Dry nitrogen is introduced into the cavity by a plastic tube shown immediately to the left of the mirror box.

FIG. 1.

Cavity mirror in the mirror box without its cover. The mirrors outside the mirror box form the combining optics. The reflectivity of the cavity mirror is greater than 99.7%. Small amounts of light from the counter-propagating laser beams are transmitted through the dielectric cavity mirror and directed by the mirrors outside the viewports onto the beam splitter. The vertical adjustor is part of the beam splitter. When the transmitted light is collimated and focused by the bottom mirror on the photodiode at the extreme left of the breadboard, a beat frequency proportional to the rotation rate of the cavity is created. Dry nitrogen is introduced into the cavity by a plastic tube shown immediately to the left of the mirror box.

Close modal
FIG. 2.

One of the Brewster windows on the plasma tube. The Brewster windows polarize the laser beams. The wire taking the high voltage to the anode is shown protruding through the aperture on the right side of the plasma tube container. To seal the container, the anode wire is sandwiched in the groove between two short slightly tapered Delran cylinders that are forced into the aperture. The bottom cylinder is shown at the extreme upper right of the figure. The negative terminal of the high voltage power supply is connected to the aluminum plasma tube container. The small wire shown in the figure connects the plasma tube container to the cathode of the plasma tube. Cables from the high voltage power supply are not shown. The plasma tube excites two counter-propagating beams in the cavity. To achieve single longitudinal mode operation, the laser operates near threshold at a current between 5 and 7 mA.

FIG. 2.

One of the Brewster windows on the plasma tube. The Brewster windows polarize the laser beams. The wire taking the high voltage to the anode is shown protruding through the aperture on the right side of the plasma tube container. To seal the container, the anode wire is sandwiched in the groove between two short slightly tapered Delran cylinders that are forced into the aperture. The bottom cylinder is shown at the extreme upper right of the figure. The negative terminal of the high voltage power supply is connected to the aluminum plasma tube container. The small wire shown in the figure connects the plasma tube container to the cathode of the plasma tube. Cables from the high voltage power supply are not shown. The plasma tube excites two counter-propagating beams in the cavity. To achieve single longitudinal mode operation, the laser operates near threshold at a current between 5 and 7 mA.

Close modal

An examination of the path of the Mayflower-Vilonia tornado shown in Figure 3 provides the tornado's relation to the ring laser in Conway, Arkansas, and the Doppler radar in Little Rock, Arkansas. The tornado was on the ground from ∼7:06 p.m. to 8:02 p.m. CDT18 on April 27, 2014 (0006-0059 UTC on April 28). At ∼7:34 p.m. CDT on April 27, 2014, the tornado destroyed sections of Mayflower with damage rated at an EF4. At ∼7:50 p.m. CDT, it entered Vilonia, Arkansas, and destroyed large sections of the town with a damage rating of EF4. At ∼8:02 p.m. CDT, the parent storm lifted. At the closest approach, it came within ∼21 km of the ring laser. The location of the tornado was updated approximately every five minutes by radar scans. Descriptions of the tornado and the EF rating were supplied by the National Weather Service (NWS) in Little Rock, Arkansas,18 plus an after-the-fact damage survey19 that was provided by Marshall et al. An FFT of the infrasound detected by the ring laser over the 56 min life of the storm yielded a coherent frequency of ∼0.94 Hz. The 0.94 Hz frequency was detected 30 min before the funnel was reported on the ground; it continued until the funnel lifted. However, the long integration times required to develop FFT peaks from the incoherent noise may mask short term frequency variations.

FIG. 3.

April 27, 2014, Mayflower-Vilonia, Arkansas EF4 tornado track of 66 km is shown in the modified image initially provided by the Little Rock, Arkansas, National Weather Service. The storm touched down in Central Arkansas at ∼7:06 p.m. CDT on April 27, 2014. At ∼7:34 p.m. CDT, the tornado destroyed sections of Mayflower, Arkansas. At ∼7:50 p.m. the tornado entered Vilonia, Arkansas. Shortly before 8:00 p.m. CDT, the tornado swept dozens of homes down to their concrete slabs in a Vilonia, Arkansas, subdivision. After approximately 56 min on the ground, the parent vortex began to lift. Tornadic infrasound was detected by a ring laser interferometer in Conway, Arkansas, and the tornado was tracked by the NWS Doppler radar in Little Rock.

FIG. 3.

April 27, 2014, Mayflower-Vilonia, Arkansas EF4 tornado track of 66 km is shown in the modified image initially provided by the Little Rock, Arkansas, National Weather Service. The storm touched down in Central Arkansas at ∼7:06 p.m. CDT on April 27, 2014. At ∼7:34 p.m. CDT, the tornado destroyed sections of Mayflower, Arkansas. At ∼7:50 p.m. the tornado entered Vilonia, Arkansas. Shortly before 8:00 p.m. CDT, the tornado swept dozens of homes down to their concrete slabs in a Vilonia, Arkansas, subdivision. After approximately 56 min on the ground, the parent vortex began to lift. Tornadic infrasound was detected by a ring laser interferometer in Conway, Arkansas, and the tornado was tracked by the NWS Doppler radar in Little Rock.

Close modal

The deployment location of the ring laser in the basement of a multistory building is not ideal. Wind loading of the building is a significant issue. Thunder easily couples into the ring laser, and frequency fluctuations are introduced by local noise. To remove outlying points due to building movement or large incoherent local noise, an upper and lower bound of 377–373 Hz was set up centered on the carrier (Sagnac) frequency in Figure 4. A program automatically eliminated points outside these bounds. The removed points were replaced by the average of the first valid points on either side of the removed entries.

FIG. 4.

Top graph in the figure shows the frequency modulation of the carrier frequency during the nearly 56 min (19:06–20:02 p.m. CDT April 27, 2014) that the EF4 Mayflower Vilonia tornado was on the ground. The graph started at 19:00 p.m. and the tornado was reported on the ground at 19:06 p.m. The breaks in the data occurred during strong wind gusts that shook the building and the basement slab on which the ring laser was mounted. The bottom graph shows an FFT of the data shown in the top graph. The FFT peak is at 0.94 Hz. The long integrating time may obscure short term frequency variations in the tornadic infrasound. The 0.94 Hz frequency was initially detected 30 min before the funnel was reported on the ground. It is not possible from the infrasound record to conclusively identify when the tornado touched down.

FIG. 4.

Top graph in the figure shows the frequency modulation of the carrier frequency during the nearly 56 min (19:06–20:02 p.m. CDT April 27, 2014) that the EF4 Mayflower Vilonia tornado was on the ground. The graph started at 19:00 p.m. and the tornado was reported on the ground at 19:06 p.m. The breaks in the data occurred during strong wind gusts that shook the building and the basement slab on which the ring laser was mounted. The bottom graph shows an FFT of the data shown in the top graph. The FFT peak is at 0.94 Hz. The long integrating time may obscure short term frequency variations in the tornadic infrasound. The 0.94 Hz frequency was initially detected 30 min before the funnel was reported on the ground. It is not possible from the infrasound record to conclusively identify when the tornado touched down.

Close modal

Figure 5 shows the FFTs of four representative tornadoes, including the Mayflower-Vilonia storm. The tornadoes were identified by comparing their corrected arrival times with the tornadoes listed in the NOAA Storm Events Database and NWS weather reports. Since infrasound can travel long distances with limited attenuation, infrasound from numerous tornadoes can sometimes arrive at close to the same time. Consequently, there is a potential uncertainty in the identity of which storms actually created the infrasound response. This is particularly true for the infrasound peaks in Figure 5(d) created from the June 4, 2015, tornado outbreak in Colorado. The amplitudes of the FFT graphs provide no information about the intensity between storms.

FIG. 5.

FFTs of four tornadoes listed by the date of detection are shown in the figure. Tornadoes create a range of signals including coherent acoustic infrasound. After corrections for travel time, coherent infrasound from all four tornadoes was detected at least 30 min before the funnel was reported on the ground and infrasound continued while the storms were on the ground. Infrasound ceased after the funnel lifted. The FFT amplitudes are relative for the tornado in each column. However, the FFT amplitudes of different storms are independent of each other. The top row shows the FFT peaks that were detected starting 30 min before the funnels were reported on the ground. Infrasound continued while each tornado was on the ground as shown in the second row of FFTs. Coherent infrasound responses ceased shortly after the funnels lifted as shown in the bottom row of FFTs. Tornado (a) was a magnitude EF3 tornado observed on April 25, 2011, in Garland County, Arkansas; tornado (b) was the EF4 tornado in Mayflower and Vilonia, Arkansas (Pulaski and Faulkner Counties), observed on April 27, 2014; tornado (c) was an EF2 storm observed in Howard County, Arkansas, on May 10, 2015. The first FFT tornado peak in (d) was apparently from an EF3 tornado that touched down on the Boulder and Larimer County line in Colorado on June 4, 2015. The source of the second FFT in (d) is less certain. It is probably from a swarm of EF0 storms also in Colorado during the same time period as the EF3 storm.

FIG. 5.

FFTs of four tornadoes listed by the date of detection are shown in the figure. Tornadoes create a range of signals including coherent acoustic infrasound. After corrections for travel time, coherent infrasound from all four tornadoes was detected at least 30 min before the funnel was reported on the ground and infrasound continued while the storms were on the ground. Infrasound ceased after the funnel lifted. The FFT amplitudes are relative for the tornado in each column. However, the FFT amplitudes of different storms are independent of each other. The top row shows the FFT peaks that were detected starting 30 min before the funnels were reported on the ground. Infrasound continued while each tornado was on the ground as shown in the second row of FFTs. Coherent infrasound responses ceased shortly after the funnels lifted as shown in the bottom row of FFTs. Tornado (a) was a magnitude EF3 tornado observed on April 25, 2011, in Garland County, Arkansas; tornado (b) was the EF4 tornado in Mayflower and Vilonia, Arkansas (Pulaski and Faulkner Counties), observed on April 27, 2014; tornado (c) was an EF2 storm observed in Howard County, Arkansas, on May 10, 2015. The first FFT tornado peak in (d) was apparently from an EF3 tornado that touched down on the Boulder and Larimer County line in Colorado on June 4, 2015. The source of the second FFT in (d) is less certain. It is probably from a swarm of EF0 storms also in Colorado during the same time period as the EF3 storm.

Close modal

Table I summarizes the results from the NOAA Storm Events Data Center and the Little Rock NWS that are pertinent to Figure 5. Comments on the tornados are provided after the table.

TABLE I.

Results from the Storm Events Data Center pertinent to Figure 4.

LocationDateStart timeEnd timeEF scaleFrequency (Hz)Width (m)Path length (km)
Garland, AR 4/25/11 17:07 CST 17:16 CST EF3 2.0 274 8.74 
Faulkner, AR 4/27/14 19:06 CDT 20:02 CDT EF4 0.94 1207 66 
Howard, AR 5/10/15 22:18 CST 22:34 CST EF2 2.25 366 10.8 
Boulder/Larimer, CO 6/4/15 17:30 MST 18:08 MST EF3 2.41 402 9.7 
LocationDateStart timeEnd timeEF scaleFrequency (Hz)Width (m)Path length (km)
Garland, AR 4/25/11 17:07 CST 17:16 CST EF3 2.0 274 8.74 
Faulkner, AR 4/27/14 19:06 CDT 20:02 CDT EF4 0.94 1207 66 
Howard, AR 5/10/15 22:18 CST 22:34 CST EF2 2.25 366 10.8 
Boulder/Larimer, CO 6/4/15 17:30 MST 18:08 MST EF3 2.41 402 9.7 

Long integration times were needed to allow the FFT routines time to extract clear coherent tornadic infrasound peaks. There were 21 tornadoes listed in the NOAA Storms Event Database for June 4, 2015, in Colorado, between 15:38 and 18:10 MST ranging in intensity from EF0 to EF3. The ring laser 2.41 Hz response in Figure 5(d) correlated best with the EF3 tornado that was located along the border of Boulder and Larimer counties in Colorado. Ten EF0 tornadoes were also listed during approximately the same time that the 2.41 Hz tornado was active. The 3.11 Hz response in Figure 5(d) is thought to have been generated by an EF0 tornado that was on the ground from 17:21 to 17:33 MST in Elbert County. Because of uncertainties in identifying the EF0 storm, it is not listed in Table I. The responses of 2.41 and 3.11 Hz in Figure 5(d) show the ability of the ring laser to distinguish between tornadoes with different funnel diameters that arrive at approximately the same time.

Assuming that R represents the funnel radius, Abdullah's theoretical model7 as simplified by Bedard6 suggests that the relation between the tornadic fundamental infrasound frequency and funnel radius is 207/R, and the relation between the first harmonic frequency and funnel radius is 371/R. Also, Abdullah's model suggests that there should be an inverse relation between the detected infrasound frequency and funnel diameter. Table II is arranged by increasing funnel diameter and shows the ring laser measured infrasound frequency versus Abdullah's theoretical values for the fundamental and first harmonic as a function of vortex radius.

TABLE II.

Laser measured infrasound versus Abdullah's theoretical fundamental and first harmonic frequencies.

LocationDateStart timeEnd timeWidth (m)Measured (Hz)Fundamental (Hz)1st harmonic (Hz)
Garland, AR 4/25/11 17:07 CST 17:16 CST 274 2.0 1.51 2.70 
Howard, AR 5/10/15 22:18 CST 22:34 CST 366 2.25 1.13 2.03 
Boulder/Larimer, CO 6/4/15 17:30 MST 18:08 MST 402 2.41 1.03 1.85 
Faulkner, AR 4/27/14 19:06 CDT 18:08 MST 1207 0.94 0.343 0.61 
LocationDateStart timeEnd timeWidth (m)Measured (Hz)Fundamental (Hz)1st harmonic (Hz)
Garland, AR 4/25/11 17:07 CST 17:16 CST 274 2.0 1.51 2.70 
Howard, AR 5/10/15 22:18 CST 22:34 CST 366 2.25 1.13 2.03 
Boulder/Larimer, CO 6/4/15 17:30 MST 18:08 MST 402 2.41 1.03 1.85 
Faulkner, AR 4/27/14 19:06 CDT 18:08 MST 1207 0.94 0.343 0.61 

Since there is a disagreement in the literature concerning the theoretical underpinning of Abdullah's formulation, some caution is needed in comparing the results in Table II. In a 2010 paper, Schmitter20 presents results from using a model confined to the lower part of the tornado vortex to study sources and possible generating mechanisms of sound. His results suggest that significant differences occur in the relation between the emitted frequencies and vortex radius depending on the boundary conditions. Under certain boundary conditions and assumptions, his results are in agreement with Abdullah's analytically derived results. On the other hand, Schecter21 in a 2012 paper suggests there are mis-steps in Abdullah's derivations that cast doubt on its fundamental credibility. In addition, he states that the principal axisymmetric oscillations of a subsonic Rankine vortex (axisymmetric Kelvin modes) do not emit acoustic radiation. In 2005, Hurricane Wilma14 experienced an eyewall replacement. A satellite image clearly showed the inner eyewall and the outer eyewall in the hurricane. The acoustic-gravity wave frequencies measured by the ring laser were inversely related to the two radii.

Figure 6 shows the long distance over which the ring laser can detect the coherent acoustic-gravity waves.

FIG. 6.

Low frequency coherent acoustic-gravity waves can travel thousands of miles with minimum attenuation. The ring laser detected acoustic-gravity waves of ∼7.1 mHz when Super Typhoon Haiyan came ashore in the Philippine Islands on November 7, 2013, at ∼20:40 UTC. Due to the travel time, the infrasound was not received in Arkansas until November 8, 2013, starting at ∼0704 UTC. Simultaneously, a ∼4.2 mHz response from the explosive eruptions of Volcano Shiveluch in Eastern Russia was also detected by the ring laser on November 8, 2013. The cluster of small peaks between approximately 3.0 and 3.3 mHz is in the buoyancy frequency regime. The restoring force for buoyancy waves is gravity; they can be created by numerous sources such as cumulous convection, flow over topography, and fluid instabilities.

FIG. 6.

Low frequency coherent acoustic-gravity waves can travel thousands of miles with minimum attenuation. The ring laser detected acoustic-gravity waves of ∼7.1 mHz when Super Typhoon Haiyan came ashore in the Philippine Islands on November 7, 2013, at ∼20:40 UTC. Due to the travel time, the infrasound was not received in Arkansas until November 8, 2013, starting at ∼0704 UTC. Simultaneously, a ∼4.2 mHz response from the explosive eruptions of Volcano Shiveluch in Eastern Russia was also detected by the ring laser on November 8, 2013. The cluster of small peaks between approximately 3.0 and 3.3 mHz is in the buoyancy frequency regime. The restoring force for buoyancy waves is gravity; they can be created by numerous sources such as cumulous convection, flow over topography, and fluid instabilities.

Close modal

The ∼7.1 mHz FFT peak is from the acoustic-gravity waves generated by Super Typhoon Haiyan as it made landfall in the Philippine Islands on November 7, 2013, at ∼20:40 UTC.22 Due to travel time, the acoustic-gravity waves were not received in Arkansas until November 8 between ∼0704 and 0834 UTC. Similar ∼7.1 mHz acoustic-gravity waves were detected by the ring laser in 2005 when Hurricanes Katrina and Wilma made landfall14 in the Gulf of Mexico. The hurricane acoustic-gravity waves were also detected in 2007 when Hurricane Dean passed between the islands of Martinique and St. Lucia in the Caribbean.14 Apparently, the peak at ∼4.2 mHz is from the explosive eruption of Volcano Shiveluch in eastern Russia. It arrived simultaneously with the waves from Super Typhoon Haiyan. The Smithsonian/USGS Global Volcanism Weekly Volcanic Activity Report for November 1–8, 2013, listed strong explosive eruptions from Sheveluch.23 Using microbarographs, Tahira et al. recorded air waves in Japan of ∼4.4 mHz during the explosive eruption of Mount Pinatubo24 in the Philippines on June 15, 1991. Volcanic infrasound peaks between 3.7 and 5.6 mHz were detected in Sweden when Mt. Saint Helens erupted in Washington State25 on May 18, 1980. The cluster of frequencies between ∼3.0 and 3.3 mHz is in the buoyancy wave regime. Their source is unknown. Buoyancy wave periods range from 5 to 120 min. The restoring force is gravity; they have several sources such as cumulous convection, flow over topography, and fluid instabilities. Typical buoyancy waves have vertical wavelengths from 5 to 15 km, horizontal phase speeds of up to 80 m/s, and a horizontal wavelength from 10 to 200 km.26 In a 1998 theoretical paper, Lognonne et al. developed a model27 predicting that coupling between the ground and atmosphere occurs at a set of frequencies related to fundamental and harmonics of atmospheric normal modes. These modes were calculated as 3.681, 4.405, 4.696, 5.076, 6.104, 7.067, 8.118, and 9.171 mHz. Responses close to these atmospheric modes have been detected by the ring laser from volcanoes or hurricanes/typhoons. Specifically, the 7.067 mHz mode is very close to the response from Super Typhoon Haiyan and hurricanes in the Gulf of Mexico as they made landfall. Frequencies from explosive volcanic eruptions have been detected by the ring laser at 3.7, 4.4, 4.7, 5.0, and 6.2 mHz. Frequencies of 7.1, 8.1, and 9.2 mHz were detected in association with tropical and extratropical hurricanes.

The ring laser interferometer is a very broad band instrument. Assuming sufficient coherence, FFTs can be used to reveal the frequency signatures from multiple sources that arrive simultaneously. The book28 by Gossard and Hooke provides a good reference and a comprehensive review of the generation and propagation of atmospheric waves including a chapter devoted to infrasound. Various types of waves can exist in the atmosphere. The acoustic-gravity waves travel at the speed of sound and their restoring force is compression. When the wavelengths of acoustic waves become long enough, gravity becomes a factor in the restoring force.5 Lamb waves29 are compressional atmospheric free oscillations that can be excited by impulsive forces such as volcanic eruptions; they propagate horizontally along the surface of the earth at the speed of sound. Their energy is mostly in the troposphere.

Laser interferometers have numerous configurations and uses. For example, in a relatively early paper, Herskovitz30 describes the use of a Michelson interferometer as an interferometric manometer. Michelson interferometers use a beam splitter to produce two laser beams that can be recombined to create an interference pattern. Gas pressure changes in one of the beams change its optical path length. Consequently, the fringes in the interference pattern shift and allow pressure changes to be measured. An active cavity such as the one used in the ring laser interferometer presented in the current paper has the excitation source (plasma tube) inside the cavity. It produces a frequency change due to pressure induced perturbations inside the cavity. It is much easier to measure small frequency changes as opposed to fringe shifts.

Ideally, the counter-propagating waves in a ring laser rotation detector are independent oscillations; however, in practice, there is cross coupling introduced by optical back scatter2–4 from one counter-propagating wave into the other. If the cross coupling is sufficiently strong, the two counter-propagating waves are injection-locked to the same frequency, which causes the beat note to disappear. Lock-in2–4 is a ubiquitous characteristic of all coupled oscillators, including ring lasers. Apparently, lock-in was first observed by Christiaan Huygens31 in 1665, when swings of the pendulums in two clocks mounted on the same wall became perfectly synchronized. Siegman32 devotes a chapter of his book in a very lucid discussion of injection locking in general and how it leads to ring laser lock-in. The cavity perimeter of 22.6 m in the Arkansas laser is large enough to allow Earth's rotation to bias the ring laser out of lock-in. However, some residual coupling remains. Back scatter induced residual coupling can create frequency pulling2–4 between the counter propagating waves. If the backscatter increases, the frequency pulling between the counter-propagating waves slightly increases. Therefore, the carrier (beat) frequency will slightly decrease. If the scatter decreases, the carrier frequency will slightly increase. Consequently, the 375 Hz carrier frequency is frequency modulated. Optical elements such as plasma tubes are not generally used in ring lasers. A plasma tube in the beam line increases backscatter and can introduce stress birefringence.2 Since acoustic waves tend to be more coherent than non-acoustic vibrations,6 FFTs can be used to extract the frequency spectrum of the coherent signals.

To our knowledge, the 16 m2 ring laser4 at the Fundamental Station in Wettzell, (Bavaria) Germany, is the most sensitive and stable ring laser currently in operation. Its lock-in threshold is cited as 0.010 Hz.4 This incredibly low value can be attributed to superb vault and cavity engineering plus improvements in laser mirror technology resulting in the so-called “super mirrors.” Variations in the scale factor (4A/λP), in the sensor orientation, and in Earth's rotation influence the frequency in Sagnac interferometers. Even in the best designs like the ring laser4 in Germany, atmospheric pressure loading of the deployment site, microseismic activity, and solid Earth tides can affect the Sagnac frequency. In the Arkansas interferometer, variations in Earth's rotation are too small to be detected; however, variations in the scale factor can result from changes in the laser cavity geometry and backscatter coupling. The changes in the cavity geometry and sensor orientation can result from distortions of the basement slab in our three story science building. The basement slab distortions can result from atmospheric pressure loading of the deployment site, wind loading of the building, and uneven solar heating. In addition to these effects, in the Arkansas laser, pressure variations can couple into the nitrogen filled cavity and introduce minute perturbations in the 22.6 m cavity. These pressure variations can slightly perturb the cavity mirrors, cause minute changes in the effective cavity length, and slightly perturb the plasma tube. All these effects can vary the backscatter which in turn modulates the carrier frequency by small amounts.

A number of years ago the corresponding author conducted a series of experiments examining lock-in thresholds. An equilateral triangular ring laser with a perimeter of 150 cm was placed on a turntable. The laser employed a plasma tube, industrial quality dielectric mirrors, and a cavity filled with air. Distortion of the sine wave from the ring laser due to frequency pulling became clearly observable in an oscilloscope below approximately 1.5 kHz. The frequency pulling became increasingly nonlinear as the lock-in threshold was approached. In a commercial navigational ring laser with a perimeter of 40 cm, using 1970 vintage mirror technology, without a plasma tube, and sealed from the atmosphere, Aronowitz2 cited the lock-in threshold as approximately 300 Hz. High levels of back scatter create significant coupling between the counter-propagating waves; however, the size of the Arkansas ring laser provides enough bias to keep it out of lock-in. In summary, our results suggest that one source of backscatter modulation is infrasound introduced perturbations of the laser cavity. In turn, frequency pulling due to backscatter changes can modulate the 375 Hz carrier frequency. FFTs can then extract the coherent infrasound and acoustic-gravity wave signatures from the non-coherent pressure and noise perturbations.

Although the coupling mechanism between coherent atmospheric infrasound and the ring laser is somewhat complex, the validity of using the ring laser as a coherent infrasound detector was verified in a controlled experiment. An FFT of the ring laser output agreed with the driving frequency of a collocated Helmholtz acoustic generator.14 Comparison with results from various NOAA studies6–9 also suggest that the ring laser has the potential to serve as an infrasound detector. Although four tornadoes are too few to draw many conclusions, the inverse relation between tornadic infrasound frequencies and tornado funnel diameters is supported by the Mayflower-Vilonia, Arkansas (Faulkner County), tornado when compared to the other three storms in Table II. In all four tornadoes, infrasound was consistently detected before the funnel was reported on the ground as was also observed by NOAA.6,8,9

A single instrument is of limited value as a tornado detector. To be of practical value, it would require an array of ring lasers and/or companion instruments such as Doppler radars. In areas where Doppler radar coverage is limited, the detection of tornadic generated infrasound is a candidate to supplement radar coverage. As large hurricanes14 move ashore, they emit infrasound. This suggests that ring laser interferometers may have the potential to supplement Doppler radar and satellites in the study of large convective storms. For example, the ring laser's ability to detect eye wall replacements14 in hurricanes may be useful in determining if this or related mechanisms are associated with hurricane intensification. Jet engines can be destroyed by volcanic ash. The extremely long distances over which an array of ring laser interferometers could detect explosive volcanic eruptions might be useful in locating eruptions in remote regions such as the Aleutian Islands. The ring laser's sensitivity to frequencies as low as 0.1 mHz allows it to detect responses in the buoyancy wave spectrum.26 It has been suggested that buoyancy waves may play a role in the intensification of tornadoes.33 The ability to show characteristic frequency spectra may help in identifying infrasound sources. Since the ring laser can detect both seismic waves and infrasound, ring lasers may have the potential for studying ground-atmospheric coupling.

This material was based in part upon work supported by the National Science Foundation under Grant No. 1147919 RUI: Geophysical Measurements Using Ring Laser sand Arrays. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation. We also acknowledge the support of the NASA Arkansas Space Grant Consortium. A note of appreciation is due the reviewers whose comments were very helpful and significantly improved the paper.

1.
W. M.
Macek
,
D. T. M.
Davis
, Jr.
,
R. W.
Olthius
,
J. R.
Schneider
, and
G. R.
White
, “
Ring laser rotation rate sensor
,” in
Optical Lasers
, edited by
J.
Fox
(
Polytechnic
,
Brooklyn
,
1963
), pp.
199
207
.
2.
F.
Aronowitz
, “
The laser gyro
,” in
Laser Applications
, edited by
M.
Ross
(
Academic Press
,
New York
,
1971
), Vol.
1
, pp.
133
200
.
3.
W. W.
Chow
,
J.
Gea-Banaclocke
,
L. M.
Pedrotti
,
V. E.
Sanders
,
W.
Schleich
, and
M. O.
Scully
, “
The ring laser gyro
,”
Rev. Mod. Phys.
57
,
61
103
(
1985
).
4.
K. U.
Schreiber
and
J.-P. R.
Wells
, “
Invited review article: large ring lasers for rotation sensing
,”
Rev. Sci. Instrum.
84
,
041101
(
2013
).
5.
A. J.
Bedard
, Jr.
and
T. M.
Georges
, “
Atmospheric Infrasound
,”
Phys. Today
53
(
3
),
32
37
(
2000
).
6.
A. J.
Bedard
, “
Low frequency atmospheric acoustic energy associated with vortices produced by thunderstorms
,”
Mon. Weather Rev.
133
,
241
263
(
2005
).
7.
A. J.
Abdullah
, “
The musical sound emitted by a tornado
,”
Mon. Weather Rev.
94
,
213
220
(
1966
).
8.
A. J.
Bedard
, Jr.
,
B. W.
Bartram
,
R. T.
Nishiyama
,
A. N.
Keane
, and
D. C.
Welsh
, “
The Infrasound (ISNET): Background, design details, and display capabilities as an 88D adjunct tornado detection tool. Session 1.1
,” in
Proceedings of the 22nd Conference on Severe Local Storms
, Hyannis, MA, 4–8 October (
American Meteorological Society
,
Boston, MA
,
2004
).
9.
A. J.
Bedard
, Jr.
,
B. W.
Bartram
,
B.
Entwistle
,
J.
Golden
,
S.
Hodanish
,
R. M.
Jones
,
R. T.
Nishiyama
,
A. N.
Keane
,
L.
Mooney
,
M.
Nicholls
,
E. J.
Szoke
,
E.
Thaler
, and
D. C.
Welsh
, “
Session P-2.8 Overview of the ISNet data set and conclusions and recommendations from a March 2004 workshop to review ISNet data
,” in
Proceedings of the 22nd Conference on Severe Local Storms
, Hyannis, MA, 4–8 October (
American Meteorological Society
,
Boston, MA
,
2004
).
10.
A.
Pancha
,
T.
Webb
,
G.
Stedman
,
D.
McLeod
, and
K.
Schreiber
, “
Ring laser detection of rotations from teleseismic waves
,”
Geophys. Res. Lett.
27
(
21
),
3553
3556
, doi: (
2000
).
11.
H.
Igel
,
K. U.
Schreiber
,
A.
Flaws
,
B.
Schuberth
,
A.
Velikoseltsev
, and
A.
Cochard
, “
Rotational motions induced by the M8.1 Tokachi-oki earthquake, September 25, 2003
,”
Geophys. Res. Lett.
32
,
L08309
, doi: (
2005
).
12.
R. W.
Dunn
,
H. H.
Mahdi
, and
H. J.
Al-Shukri
, “
Design of a relatively inexpensive ring laser seismic detector
,”
Bull. Seismol. Soc. Am.
99
(
2B
),
1437
1442
(
2009
).
13.
K. U.
Schreiber
,
J. N.
Hautmann
,
A.
Velikoseltsev
,
J.
Wassermann
,
H.
Igel
,
J.
Otero
,
F.
Vernon
, and
J. P. R.
Wells
, “
Ring laser measurements of ground rotations for seismology
,”
Bull. Seismol. Soc. Am.
99
(
2B
),
1190
1198
(
2009
).
14.
R. W.
Dunn
,
W. V.
Slaton
, and
L. M.
Kendall
, “
Detection of low frequency hurricane emissions using a ring laser interferometer
,”
J. Appl. Phys.
112
,
073110
(
2012
).
15.
R. W.
Dunn
and
A. R.
Hosman
, “
Detection of volcanic infrasound with a ring laser interferometer
,”
J. Appl. Phys.
116
,
173109
(
2014
).
16.
F. G.
Stremler
,
Introduction to Communication Systems
, 2nd ed. (
Addison-Wesley
,
Reading, MA
,
1982
), pp.
315
326
.
17.
L. E.
Kinsler
,
A. R.
Frey
,
A. B.
Coppens
, and
J. V.
Sanders
,
Fundamentals of Acoustics
, 3rd ed. (
Wiley
,
New York
,
1982
), pp.
225
227
.
18.
See www.srh.noaa.gov/lzk/?n=svr0414c.htm for NWS Little Rock; and links of interest.
19.
T. M.
Marshall
,
J.
Robinson
, and
L.
Tanner
, “
Damage survey of the Mayflower-Vilonia, Arkansas tornado
,” in
American Meteorological Society 27th Conference on Severe Local Storms
(
2014
).
20.
E. D.
Schmitter
, “
Modelling tornado dynamics and generation of infrasound, electric and magnetic fields
,”
Nat. Hazards Earth Syst. Sci.
10
,
295
298
(
2010
).
21.
D. A.
Schecter
, “
A brief critique of a theory used to interpret the infrasound of tornadic thunderstorms
,”
Mon. Weather Rev.
140
,
2080
2089
(
2012
).
22.
23.
See Volcano.si.edu for Global Volcanism Program Smithsonian/USGS Weekly Volcanic Activity Report (last accessed November 1–8,
2013
).
24.
M.
Tahira
,
M.
Nomura
,
Y.
Sawada
, and
K.
Kamo
, “
Infrasound and acoustic-gravity waves generated by the Mount Pinatubo eruption of June 15, 1991
,” in
Fire and Mud Eruptions and Lahars of Mount Pinatubo, Philippines
(
U.S. Geological Survey Publication
,
2004
).
25.
M.
Bath
, “
Atmospheric waves from Mount St. Helens
,”
EOS
63
(
11
),
193
(
1982
).
26.
C. D.
de Groot-Hedlin
,
M. A. H.
Hedlin
, and
D. P.
Drob
, “
Atmospheric variability and infrasound monitoring
,” in
Infrasound Monitoring for Atmospheric Studies
, edited by
A.
Le Pichon
 et al. (
Springer Science
,
New York
,
2010
), pp.
490
492
.
27.
P.
Lognonne
,
E.
Clevede
, and
H.
Kanamori
, “
Computation of seismograms and atmospheric oscillations by normal-mode summation for a spherical earth model with realistic atmosphere
,”
Geophys. J. Int.
135
,
388
406
(
1998
).
28.
E. E.
Gossard
and
W. H.
Hooke
,
Waves in the Atmosphere: Atmospheric Infrasound and Gravity Waves, their Generation and Propagation
(
Elsevier
,
New York
,
1975
).
29.
R. S.
Lindzen
and
D.
Blake
, “
Lamb waves in the presence of realistic distributions of temperature and dissipation
,”
J. Geophys. Res.
77
(
12
),
2166
, doi: (
1972
).
30.
S. B.
Herskovitz
, “
Simple interferometric manometer
,”
Rev. Sci. Instrum.
37
,
452
(
1966
).
31.
P. S.
Spoor
and
G. W.
Swift
, “
The Huygens entrainment phenomenon and thermos-acoustic engines
,”
J. Acoust. Soc. Am.
108
(
2
),
588
(
2000
).
32.
A. E.
Siegman
,
Lasers
(
University Science Books
,
Sausalito, CA
,
1986
), pp.
1129
1170
.
33.
See http//science1.nasa.gov/science-news/science-at-nasa/2008/19mar_grits/ for how buoyancy waves can influence tornadoes.