Passive acoustic inversion techniques for measuring gas flux into the water column have the potential to be a powerful tool for the long-term monitoring and quantification of natural marine seeps and anthropogenic emissions. Prior inversion techniques have had limited precision due to lack of constraints on the initial amplitude of a bubble's excitation following its release into the water column (Rε0i). Rε0i is determined by observing the acoustic signal of bubbles released from sediment in a controlled experiment and its use is demonstrated by quantifying the flux from a volcanic CO2 seep offshore Panarea (Italy), improving the precision by 78%.

In recent years, there has been a growing need to refine methods for quantifying the volume of gas released into the water column from marine sediments. This demand is fueled by a need to better understand the natural carbon cycle and a requirement to measure, monitor, and verify offshore carbon capture and storage (CCS) sites.1–6 While there are a variety of techniques to quantify gas emissions, including direct physical measurements, many of these are severely limited in terms of applicability in regard to spatial and temporal coverage, as well as cost. Optical techniques are generally accurate for a limited bubble size range (bubbles large enough to be accurately resolved) but limited in the area they can survey and dependent on water visibility.7–9 Active acoustic techniques, although able to cover a large area, require time-consuming processing to provide accurate quantification data and give only a snapshot measurement.10–15 Furthermore, the accuracy of the assumptions that underpin the conversion of active acoustic data into gas fluxes is still evolving.10 Passive acoustic inversion techniques, on the other hand, are low energy and capable of long-term deployments in any level of visibility and also provide information about the size distribution of bubbles released from the seep.7,16–21

Passive acoustic inversion relies on a physical model of the sound emitted by a bubble as it is formed. Originally, this principle was restricted to measuring the size of individual bubbles in low flux environments, via the so-called signature method, to ensure the acoustic signature of bubbles did not overlap.22–25 This approach relies on knowing the frequency emitted by a bubble immediately after release, which is directly related to its equilibrium radius.19,24,26,27

Signal processing methods were developed to allow this technique to extend to intermediate flux rates when bubble signatures begin to overlap.28 For high flux rates, when bubble signatures do overlap to a significant degree, an approach based on the spectral density of the acoustic pressure measured in the far field at a known distance from a gas seep is appropriate.21,29 Using a model of the spectrum of the signature of each bubble formation event, the measured spectrum can be used to infer the bubble size distribution. However, the energy released by an individual bubble remains poorly characterized,21 particularly, how it varies with depth and mode of injection. This quantity directly affects the gas flux estimate, for example, if the model assumes the bubble is a stronger sound source than it is, then the inversion will underestimate the gas flux.21 Conversely, if a single bubble emits more energy than predicted by the model, then the passive acoustic inversion will overestimate the gas flux. To date, the use of the spectral method in the field has proven effective, providing continuous estimates of gas flux over extended periods of time validated by intermittent physical and optical measurements. Unfortunately, such estimates have contained large uncertainties.7,19,30

This study examines the acoustic signature of bubbles released from coarse-grained sediment to improve the spectral inversion approach by examining the energy released by an individual bubble and the relationship between the initial radius of excitation of a bubble and the equilibrium radius of a bubble. It should be noted at this stage, while a bubble released into the water column may not be spherical, the use of a radial displacement as a descriptor is common in literature as a proxy for more complex volumetric descriptions as (a) the zeroth-order spherical harmonic displacement was the only one that changed bubble volume (to first order) and so radiated a spherical pressure wave,27 and (b) because this aligned the formulation with the historical use in the West of representing bubble volume by an equivalent spherical bubble of known radius.31 We then use this refined acoustic inversion technique to quantify ebullition from a natural CO2 seep occurring at a site with coarse-grained sediment and compare the estimate to physical and simultaneous optical measurements.

The following summarizes the mathematical model for the sound from a plume, which forms the basis of the inversion approach. The theory assumes that the acoustic pressure signature from a single bubble is an exponentially decaying sinusoid.21 

For a spherical bubble, immediately after being entrained into the water column, a bubble will undergo damped simple harmonic motion. The radius of a bubble, R, over time, t, can be described by

(1)

where Re denotes the real part of a complex number, δtot is the total damping coefficient, H is the Heaviside step function, ti is the moment when the acoustic signal is first detected, tϑ accounts for the propagation time between perturbations of the bubble and the corresponding pressure signal at the receiver, titϑ is the moment when the bubble begins to oscillate, R0 is the equilibrium radius of the bubble, and Rε0i is the initial amplitude of displacement of the bubble wall.21 

The natural angular frequency, ω0, of the bubble is given by an extended form of Minnaert's equation,19,27

(2)

where κ is the polytropic index of the gas in the bubble, p0 is the ambient pressure outside of the bubble, pv is the gas' vapour pressure, ρ0 is the density of surrounding liquid, and η is the liquid's shear viscosity.

The oscillating bubble wall creates an acoustic pressure in the far field at a distance, r, given by

(3)

The squared magnitude of the Fourier transform of the pressure radiated from a single bubble at a distance, r, is

(4)

The spectrum, S(ω), of the measured acoustic pressure from a population of bubbles with size distribution D(R0), which is defined such that R1R2D(R0)dR0 represents the number of bubbles generated per second with a radius in the range (R1,R2) can be expressed as

(5)

To estimate the number of bubbles via the far field acoustic signal, one needs to determine what value of D(R0) generates the measured power spectrum, S(ω). This can be achieved by discretising Eq. (5) into Nb finite radii bins such that the nth radius bin is centered on the radius, Rnc. Assuming that the bins are contiguous, the power spectrum can be given as

(6)

where Ψ(n) is the bubble generation rate within a radius bin (i.e., the number of bubbles formed per second within the nth radius bin) such that Ψn=Rn+1cRncD(Rnc). The spectrum can be evaluated at the Nf discrete frequencies, ωm, and organized into a column vector, S¯, allowing Eq. (6) to be written in matrix form,

(7)

where Ψ¯ is a column vector containing the elements Ψ(n), and M is an Nf×Nb matrix with elements Mm,n=X(ωm,Rnc)2.

The matrix equation (7) can be solved for the number of bubbles generated per second Ψ¯ in a radius bin, provided that S¯ and M are known, to yield an estimate of the gas flux. However, while S¯ can easily be determined accurately with the use of a calibrated hydrophone, M remains difficult to accurately determine. This is because Eq. (4) contains one important unknown value, the initial amplitude of the displacement of the bubble wall when it begins to oscillate, Rε0i. Most work on studying this has assumed a model of the form, Rε0i=χradR0, where χrad is a constant, representing the ratio of the initial amplitude of displacement of the bubble radius and equilibrium radius. The earliest estimate of χrad for bubbles of radii close to 0.001 mm measured near a small waterfall suggested that24χrad=105, and the same data were reanalyzed using a more sophisticated model to yield32χrad=104. And, considering data from smaller bubbles (R0 0.000 312 m) generated in breaking waves suggested33 that χrad=1.5×102. Generating bubbles via sheared flows allowed the number of bubbles (>1000) and the range of radii (R0= 0.0001–0.001 m) measured to be greatly expanded.34 This work suggested that χrad was not a constant and increased with decreasing bubble size. Further, in this data, χrad was seen to vary by up to 2 orders of magnitude. One approach to estimating χrad is to compute the flux of bubbles in an experiment where the flux rate can be controlled and measured directly; such an approach has been found to yield estimates of χrad between191.4×104 and 5.6×104. The apparent variable nature of χrad across bubble sizes and release mechanisms potentially introduces considerable uncertainty into derived flux estimates.19 

None of these estimates for the initial amplitude are based on bubbles emerging from sediments. Further, these estimated values are based on a small set of environmental conditions, for example, they are the range of bubble sizes and generation mechanisms.19 Consequently, there is a need for calibrated Rε0i values, particularly in conditions replicating those found in the field. This paper undertakes such an exercise.

The experimental evidence supporting the assumption that χrad is a constant is rather weak. Herein, the general approach based on assuming Rε0i is a function of R0, i.e., Rε0i=f(R0), is adopted so that Eq. (4) becomes

(8)

In this study, we analyze the acoustic emissions from 368 bubbles released from sediment into the water column to determine the Rε0i, measuring the frequency of each bubble oscillation alongside the maximum radius of excitation and total energy released.

To measure and analyze the acoustic signature of bubbles released from sediment, air was injected into the base of ∼0.30 m of artificial sediment with resulting emissions into the water column recorded with a calibrated hydrophone [Reson TC 4032, Teledyne Technology, Alton, with a sensitivity of 167 dB re 1 V/μPa at 1 m; 0.30 m above the sediment surface; Figs. 1(a) and 1(b)]. Synthetic sediment (Ballotini; 0.009 m diameter, pebbly gravel on the Wentworth scale) was used to ensure uniformity between the grains and simplify porosity and permeability calculations [Fig. 1(c)]. The experiment was designed so that ebullition was in the center of a large 8 m × 8 m (× 5 m deep) freshwater tank [Fig. 1(a)] to reduce the interference of wall reflections, allowing accurate characterization of the acoustic signatures [Fig. 1(d)]. The gas injection rate at the base of the sediment was adjusted until gas migration through the sediment caused bubbles to enter the water column at less than 1 bubble per second. Bubble radii were determined directly from optical measurements (Sony FDR3000 camera, Minato, Japan; 0.50 m from ebullition site).

FIG. 1.

(Color online) Experimental design. (a) A schematic showing gas injection at base of sediments and the position of optical and acoustic sensors within the water tank; (b) detailed schematic; (c) image of the Ballotini sediment used for the experiment; and (d) sonogram displaying the typical acoustic emission of a bubble as recorded by the hydrophone.

FIG. 1.

(Color online) Experimental design. (a) A schematic showing gas injection at base of sediments and the position of optical and acoustic sensors within the water tank; (b) detailed schematic; (c) image of the Ballotini sediment used for the experiment; and (d) sonogram displaying the typical acoustic emission of a bubble as recorded by the hydrophone.

Close modal

The optical footage was used to identify periods where single, discrete bubbles entered the water column to ensure that the acoustic signature of an individual bubble release was analyzed. The acoustic and optical data for each bubble were jointly analyzed to determine the start and end time of the bubble oscillation, as well as the maximum pressure of the bubble signal (Pmax). The initial excitation radius of the bubble (Rε0i) is then calculated using the pressure signals via 21,34

(9)

where Pmax is the maximum pressure signal recorded at the hydrophone as a result of the bubble oscillation.

The center frequency of each bubble signature was determined by iteratively fitting an exponential decaying sinusoid to the signal based on least squares error criterion. The frequency of this wave was used to calculate the equilibrium bubble radius, R0, using Eq. (2). The total energy of each bubble oscillation was determined by integrating the acoustic pressure squared as recorded at the hydrophone between the start (t1) and end (t2) of the event. This is corrected for range, assuming spherical spreading.

(10)

where c is the speed of sound in water (1500 ms−1) and fs is the sampling frequency (96 kHz). To account for background noise (electrical and acoustic), the ambient noise on the measurement was calculated using Eq. (10) with the average energy measured over 1 s being equal to 6.2 × 10−9 J. Subsequently, 6.2 × 10−9(t2t1) J were subtracted from each event so that only the energy of the bubble oscillation was observed.

To validate these observations, acoustic data collected from a natural seep in offshore Panarea (Bottaro Crater), Italy, is inverted and the estimated flux compared to that obtained by physical bottle measurements collected by a diver and estimates made from simultaneous optical measurements. Panarea is a small Aeolian Island in the southern Tyrrhenian Sea, ∼20 km SW of the active volcano Stromboli (Fig. 2). As a result of the underlying magma chamber, the waters surrounding the island are host to numerous natural CO2 seeps.35–38 The calm clear shallow waters surrounding Panarea make it an ideal natural laboratory for testing ebullition detection and quantification techniques.39–41 Bubbles from the target seep were released from coarse-grained sediment at a rate of approximately 5 bubbles per second, directly comparable to our tank experiment. Similarly, bubbles released from the seep ranged in size from 0.001 to 0.008 m comparable to the 0.003–0.012 m observed in the tank [Fig. 2(c)]. The seep was, however, located at a depth of 12 m, compared to the depth of 2.5 m in the experiments.

FIG. 2.

(Color online) The field work location map showing (a) experimental geometry at the seep site (Bottaro Crater, offshore Panarea, Italy) and (b) the acoustic optical lander deployed over the target seep in Bottaro Crater.

FIG. 2.

(Color online) The field work location map showing (a) experimental geometry at the seep site (Bottaro Crater, offshore Panarea, Italy) and (b) the acoustic optical lander deployed over the target seep in Bottaro Crater.

Close modal

Analysis focusses on 10 min of simultaneous optical and acoustic data acquired using an open frame lander, incorporating a calibrated hydrophone (Geospectrum M36, GTI, Dartmouth, Nova Scotia, Canada), and two inward facing Sony FDR-X3000 cameras, focussing on a single central point with scale boards positioned directly opposite. Optical data were processed using techniques described in Li et al., where the radius of each bubble was calculated from the assumed volume of revolution.7 Acoustic data were analyzed in 30 s blocks while the optics was analyzed in 15 s blocks.

In total, 368 single bubble releases were analyzed, and the results are discussed below. Natural frequencies were detected between 300 and 1000 Hz, corresponding to bubble radii between ∼0.003 and 0.012 m (Fig 3).

FIG. 3.

(Color online) The experimental results from the analysis of single bubbles showing (a) energy released by a bubbles oscillation vs its equilibrium radius R0 and (b) the initial amplitude of excitation Rε0i vs the equilibrium radius. The lines of best fit are shown in red alongside the assumed Rε0i values in Bergès et al. (Ref. 19) in green, (c) the initial amplitude of excitation of bubbles normalized by the equilibrium radius (χrad) vs the equilibrium radius from this study (black) alongside results from Deane and Stokes (Ref. 34; in blue) and other prior studies (black diamonds). Also shown is the range of bubble sizes used in Berges et al. and this study during passive acoustic flux inversion.

FIG. 3.

(Color online) The experimental results from the analysis of single bubbles showing (a) energy released by a bubbles oscillation vs its equilibrium radius R0 and (b) the initial amplitude of excitation Rε0i vs the equilibrium radius. The lines of best fit are shown in red alongside the assumed Rε0i values in Bergès et al. (Ref. 19) in green, (c) the initial amplitude of excitation of bubbles normalized by the equilibrium radius (χrad) vs the equilibrium radius from this study (black) alongside results from Deane and Stokes (Ref. 34; in blue) and other prior studies (black diamonds). Also shown is the range of bubble sizes used in Berges et al. and this study during passive acoustic flux inversion.

Close modal

It is useful to examine the energy released by a bubble oscillation as the degree of variation gives us an insight into how consistent the process is. The energy emitted during a bubble's oscillation varies across nearly 3 orders of magnitude, 10141011 J, with a mean value of 7.5×1013 J. Plotting the energy released against the equilibrium bubble radius [Fig. 3(a)], we see a large degree of scatter for R0 < 0.006 m, decreasing for R0 > 0.006 m.

The initial amplitude of excitation of a bubble ranged mainly between 107 and 106 m with a mean value of 6.0×107 m. Plotting Rε0i against the equilibrium bubble radius [Fig. 3(b)], we see a strong clustering of the data for larger bubbles (R0 > 0.006 m) but, once again, a larger degree of scatter for smaller bubbles (R0 < 0.006 m). Statistically, the total correlation between R0 and Rε0i, here, has a Pearson's coefficient of 0.26, a p-value of <0.001 with a trend line for R0 and Rε0i expressed in m, determined using total least squares regression (least absolute residual) of

(11)

which has a coefficient of determination of 0.99. While this relationship is highly effective for bubbles R0 > 0.006 m with the 10th and 90th percentiles of the estimates being within −10% and +15% of observed values. Below R0 < 0.006 m, there is a greater variability in the measurement points about the regression line with a root mean square error of 2.6×107 m compared to an average value of 6×107 m.

There is a weak but statistically significant, positive correlation between Rε0i and R0 in the results here, particularly, at larger bubble sizes. This contrasts with prior work which assumes that Rε0i=χradR0 for a constant χrad. It is important to note that the relationship developed here is an empirical one, which is not based on a physical model, and should only be applied to bubble sizes outside the range tested with care. To test the validity of this relationship for bubbles smaller than those tested here, the results are plotted alongside those of a previous study19,32,34 [Fig. 3(c)], where observed bubbles ranged in size from ∼0.0001 to 0.001 m. Note that this is not a like for like comparison as the experiments use different generation mechanisms: sheared bubbles in a hydrodynamic current at 2 m water depth and releasing bubbles from sediment. Despite the methodological differences, the trend line derived passes through the center of the data cloud for the previous experiment and is in general good agreement down to the smallest bubble sizes, suggesting that this model is applicable across a wider range of bubble sizes than just those studied herein.

Previous assumptions about Rε0i were made by Bergès et al.19 who used the 25th and 75th percentile of the Rε0i observations of Deane and Stokes34 to invert gas flux from needles and porous stones in a test tank. Comparing our estimates of Rε0i to those used by Bergès et al. when inverting gas flux in Fig. 3(b), we note that their assumption is consistent with our Rε0i observation between R0= ∼0.0001 and 0.003 m, where this is below the range of most bubble sizes observed here. This is consistent with the previous results, only suggesting that the error bounds provided were more conservative than necessary. However, as (1) bubbles were released via a different mechanism, meaning that Rε0i can be expected to vary slightly; and (2) the bubble sizes ranged between 0.001 and 0.004 m radius, meaning that the majority of Rε0i calculations were consistent with this study [Fig. 3(c)].

This highlights the need for researchers using acoustic flux inversion techniques in the field to compare their natural site to that used to calibrate. How similar the range of bubble sizes, flux rates, release mechanisms (needles, sediment type, etc.), depths, or even free field conditions are between the two will determine how accurate the Rε0i estimates will be. Here, an empirical linear relationship between Rε0i and R0 has been established for these specific experimental conditions (bubbles released from course sediment in fresh water at 2.5 m depth at a slow rate of a few bubbles per second). The limits of the conditions under which this relationship can reasonably predict the bubble excitation have not been established.

While our data shows a greater variability in Rε0i at smaller bubble sizes, this has little impact on the accuracy of the passive acoustic inversion technique as larger bubbles (for whom we are able to more accurately predict Rε0i) dominate the total volume of gas emitted (i.e., 1000 bubbles with a radius of 0.0006 m contribute the same volume of gas as one bubble with a radius of 0.006 m). Hence, in the field, it is more important for flux inversion purposes to accurately calculate Rε0i for large bubbles.

To test the effectiveness of the passive acoustic inversion technique based on Eq. (11), the flux from data collected in Panarea (see Sec. II B) was compared to that recorded by other techniques. A preliminary analysis of the acoustic data revealed a strong continuous signal below ∼500 Hz (Fig. 4). As optical footage suggested very few bubbles larger than 0.006 m in radius (Fig 5), f0 = 540 Hz were released, we consider this low frequency noise to likely be the result of moving sediment grains and/or distant seeps.20 Hence, the acoustic inversion is performed between 600 and 2500 Hz (corresponding to bubble radii between 0.001 and 0.005 m) to avoid interference from non-bubble related signals. The upper and lower estimates of our refined acoustic inversion were created assuming a 15% error in Rε0i based on the 95% confidence interval.

FIG. 4.

(Color online) The acoustic spectrogram of natural CO2 as recorded in Panarea during a 10 min observation period with the radius of a bubble that would be responsible for a given frequency displayed on the right axis.

FIG. 4.

(Color online) The acoustic spectrogram of natural CO2 as recorded in Panarea during a 10 min observation period with the radius of a bubble that would be responsible for a given frequency displayed on the right axis.

Close modal
FIG. 5.

(Color online) The equilibrium bubble radius distribution as estimated by each technique during a 10 min period, showing the refined acoustic inversion technique developed in this study (red), the previous acoustic inversion technique based on assuming that χRad is constant: (green) and optical inversion (blue). The upper and lower estimates of the optical measurements are shown via blue dashed lines.

FIG. 5.

(Color online) The equilibrium bubble radius distribution as estimated by each technique during a 10 min period, showing the refined acoustic inversion technique developed in this study (red), the previous acoustic inversion technique based on assuming that χRad is constant: (green) and optical inversion (blue). The upper and lower estimates of the optical measurements are shown via blue dashed lines.

Close modal

The acoustic inversion technique employing Eq. (11) produces a bubbles size distribution (modal bubble radius 0.0047 m) similar to that produced used by Bergès et al. (modal bubble radius 0.0044 m)19 with slightly more small bubbles reported by the old inversion technique (Fig. 5). Optical inversion of the Panarea data measured a single peak in the bubble size distribution at 0.0034 ± 0.0006 m.

The flux results of our refined inversion technique using Eq. (11) are presented in Fig. 6 alongside prior inversion results, assuming that χRad is between 1.4×104 and 5.6×104. Our refined inversion technique estimate of the gas flux is 3.2 ± 0.6 L/min, significantly greater than that predicted by the previous acoustic inversion technique (0.8 ± 0.7 L/min) but consistent with the mean flux value derived via optical inversion (3.0 ± 0.8 L/min). Note that the uncertainty in the method based on Eq. (11) has reduced by 78% over that employed by the previous method (reducing from 88% to 19%). However, the gas flux estimated by the refined acoustic inversion is slightly higher, 5%–8%, than that obtained by the physical measurement taken prior to deploying the lander (2.3 L/min). The failure of the prior inversion technique to match other flux estimates here is due to the larger bubble sizes seen in Panarea, since it is based on extrapolating R0= 0.002 m to larger bubbles [Fig. 3(d)].

FIG. 6.

(Color online) The gas flux estimate via each technique across the 10 min period, the refined acoustic inversion technique developed in this study (red), the previous acoustic inversion technique based on assuming that χRad is constant: (green) optical inversion and (blue) alongside the physical diver measurement taken prior to deployment (dashed black line).

FIG. 6.

(Color online) The gas flux estimate via each technique across the 10 min period, the refined acoustic inversion technique developed in this study (red), the previous acoustic inversion technique based on assuming that χRad is constant: (green) optical inversion and (blue) alongside the physical diver measurement taken prior to deployment (dashed black line).

Close modal

The overestimate in the acoustic inversion estimates of flux performed using more accurate values of Rε0i could be the result of, at least, two possible factors: (1) the role of seabed and sea surface reflections and (2) secondary bubbles oscillations after release.17,24 To address the first possibility, a model of the direct acoustic path to the hydrophone along with the path reflecting off the seafloor to the hydrophone was formed. This suggested that the energy of the reflected signal is <10% of the direct path, meaning its contribution to the overall overestimation of the flux is too small to explain the discrepancy. This would suggest that the most likely explanation for our overestimate is additional bubble oscillation occurring within the water column as bubbles merge or fragment. Wide-area video footage at Panarea shows multiple bubbles fragmenting in the water column, and this evidence is reinforced when consideration is given to the comparison of the bubble size distribution determined by optical and acoustic techniques. It may be that large bubbles are released at the seafloor where they are recorded by the hydrophone but not the optics, as they are out of thr camera's field of view. By the time the bubbles rise ∼0.75 m into the view of the camera, some have fragmented into smaller bubbles, and these smaller bubbles are measured by the optics for the first time and the acoustics for a second time, resulting in the secondary acoustic peak alongside the optical peak at ∼0.0035 m. In summary, further developments are needed in the acoustic flux inversion code to account for this complex bubble behavior.

As the precision of acoustic inversion techniques improve it will become increasingly more important to quantify the effect of secondary oscillations. To do so, researchers will need some gauge of how frequently bubbles from the seep are fragmenting or merging as compared to the rate of release. This almost certainly varies from seep to seep based on the number of factors, such as the bubble size distribution, where larger bubbles are more likely to fragment,42 and the rate of release itself, with bubbles released in rapid succession being more likely to merge together.17 One could attempt to adapt breakup frequency equations for some indication of the rate of secondary oscillation, however, these typically assume turbulent flows, unlike those seen at the seep here, and does not account for bubbles merging.43 Assuming that secondary oscillation is solely responsible for our overestimate of flux, then the resulting scale factor would be between 0.6 and 0.9.

To improve quantification of gas flux from natural seeps and anthropogenic sources by passive acoustic techniques, we have investigated the initial amplitude of the excitation of bubbles, Rε0i, released from coarse sediment at 2.5 m water depth, a previously unknown parameter in the inversion process.

We have empirically determined a relationship between Rε0i and the equilibrium radius of a bubble, R0 (in m), of the form Rε0i=9.608R0+0.447×106.

Our data show a weak positive correlation, especially at larger bubble sizes, and appears generally consistent with prior investigations of smaller (<1 mm) bubbles.

We demonstrate the potential of our refined inversion technique by inverting the flux from a natural CO2 seep in Panarea (Italy) and comparing the results to simultaneous optical measurements. Here, we find that after filtering out low frequency, non-bubble related noise, our acoustic flux estimate is consistent with optical measurements and only 5%–8% larger than prior physical measurements. We identify the post release fragmentation of bubbles as a likely explanation of minor overestimation of flux.

The 78% improvement in the precision of the acoustic inversion technique demonstrated here will allow for the more widespread adoption of passive acoustic techniques for the long-term observations of natural seeps and the measuring, monitoring, and verification (MMV) of secure CCS sites.6 

Funding was provided by the European Union's Horizon 2020 research and innovation programme under the Grant No. 654462 (Strategies for Environmental Monitoring of Marine Carbon Capture and Storage). We would like to thank Michele Deponte, Emiliano Gordini, Diego Cotterle, Valentina Esposito from the Italian National Institute of Oceanography and Applied Geophysics for their logistical support and the scientific divers Andrea Fogliozzi and Martina Gaglioti. We are grateful to the European carbon dioxide capture and storage laboratory infrastructure (ECCSEL)—NatLab Italy of Panarea, an ECCSEL-European Research Infrastructure Consortium infrastructure of Istituto Nazionale di Oceanografia e di Geofisica Sperimentale, without whom our field work would not have been possible. We thank John Davis for his contributions to the lander design and fieldwork.

1.
J.
Blackford
,
Y.
Artioli
,
J.
Clark
, and
L.
de Mora
, “
Monitoring of offshore geological carbon storage integrity: Implications of natural variability in the marine system and the assessment of anomaly detection criteria
,”
Int. J. Greenhouse Gas Control
64
,
99
112
(
2017
).
2.
M.
Razaz
,
D.
di Iorio
,
B.
Wang
,
S.
Daneshgar Asl
, and
A. M.
Thurnherr
, “
Variability of a natural hydrocarbon seep and its connection to the ocean surface
,”
Sci. Rep.
10
(
1
),
12654
(
2020
).
3.
D.
Oppo
,
L.
de Siena
, and
D. B.
Kemp
, “
A record of seafloor methane seepage across the last 150 million years
,”
Sci. Rep.
10
(
1
),
2562
(
2020
).
4.
J.
Blackford
,
H.
Stahl
,
J. M.
Bull
,
B. J. P.
Bergès
,
M.
Cevatoglu
,
A.
Lichtschlag
,
D.
Connelly
,
R. H.
James
,
J.
Kita
,
D.
Long
,
M.
Naylor
,
K.
Shitashima
,
D.
Smith
,
P.
Taylor
,
I.
Wright
,
M.
Akhurst
,
B.
Chen
,
T. M.
Gernon
,
C.
Hauton
,
M.
Hayashi
,
H.
Kaieda
,
T. G.
Leighton
,
T.
Sato
,
M. D. J.
Sayer
,
M.
Suzumura
,
K.
Tait
,
M. E.
Vardy
,
P. R.
White
, and
S.
Widdicombe
, “
Detection and impacts of leakage from sub-seafloor deep geological carbon dioxide storage
,”
Nat. Clim. Change
4
(
11
),
1011
1016
(
2014
).
5.
M.
Dean
,
J.
Blackford
,
D.
Connelly
, and
R.
Hines
, “
Insights and guidance for offshore CO2 storage monitoring based on the QICS, ETI MMV, and STEMM-CCS projects
,”
Int. J. Greenhouse Gas Control
100
,
103120
(
2020
).
6.
M.
Dean
and
O.
Tucker
, “
A risk-based framework for measurement, monitoring and verification (MMV) of the Goldeneye storage complex for the Peterhead CCS project, UK
,”
Int. J. Greenhouse Gas Control
61
,
1
15
(
2017
).
7.
J.
Li
,
P. R.
White
,
B.
Roche
,
J. M.
Bull
,
T. G.
Leighton
,
J. W.
Davis
, and
J. W.
Fone
, “
Acoustic and optical determination of bubble size distributions—Quantification of seabed gas emissions
,”
Int. J. Greenhouse Gas Control
108
,
103313
(
2021
).
8.
B.
Wang
and
S. A.
Socolofsky
, “
A deep-sea, high-speed, stereoscopic imaging system for in situ measurement of natural seep bubble and droplet characteristics
,”
Deep Sea Res., Part I
104
,
134
148
(
2015
).
9.
A.
Jordt
,
C.
Zelenka
,
J. S.
von Deimling
,
R.
Koch
, and
K.
Köser
, “
The Bubble Box: Towards an automated visual sensor for 3D analysis and characterization of marine gas release sites
,”
Sensors
15
(
12
),
30716
30735
(
2015
).
10.
J.
Li
,
B.
Roche
,
J. M.
Bull
,
P. R.
White
,
T. G.
Leighton
,
G.
Provenzano
,
M.
Dewar
, and
T. J.
Henstock
, “
Broadband acoustic inversion for gas flux auantification—Application to a methane plume at Scanner Pockmark, Central North Sea
,”
J. Geophys. Res.: Oceans
125
(
9
),
e2020JC016360
, (
2020
).
11.
J.
Schneider von Deimling
,
P.
Linke
,
M.
Schmidt
, and
G.
Rehder
, “
Ongoing methane discharge at well site 22/4b (North Sea) and discovery of a spiral vortex bubble plume motion
,”
Mar. Pet. Geol.
68
,
718
730
(
2015
).
12.
J.
Greinert
, “
Monitoring temporal variability of bubble release at seeps: The hydroacoustic swath system GasQuant
,”
J. Geophys. Res.
113
(
C7
),
C07048
, (
2008
).
13.
J.
Greinert
,
Y.
Artemov
,
V.
Egorov
,
M.
Debatist
, and
D.
McGinnis
, “
1300-m-high rising bubbles from mud volcanoes at 2080 m in the Black Sea: Hydroacoustic characteristics and temporal variability
,”
Earth Planet. Sci. Lett.
244
(
1–2
),
1
15
(
2006
).
14.
K.
Baik
,
T. G.
Leighton
, and
J.
Jiang
, “
Investigation of a method for real time quantification of gas bubbles in pipelines
,”
J. Acoust. Soc. Am.
136
(
2
),
502
513
(
2014
).
15.
T. G.
Leighton
,
K.
Baik
, and
J.
Jiang
, “
The use of acoustic inversion to estimate the bubble size distribution in pipelines
,”
Proc. R. Soc. A
468
(
2145
),
2461
2484
(
2012
).
16.
L.
Chen
,
V.
Trinh
,
W.
Yang
, and
K.
Mohanangam
, “
Prediction of bubble generation based on acoustic emission
,”
Acoust. Aust.
44
(
2
),
325
331
(
2016
).
17.
T. G.
Leighton
,
K. J.
Fagan
, and
J. E.
Field
, “
Acoustic and photographic studies of injected bubbles
,”
Eur. J. Phys.
12
(
2
),
77
85
(
1991
).
18.
J.
Li
,
P. R.
White
,
J. M.
Bull
,
T. G.
Leighton
,
B.
Roche
, and
J. W.
Davis
, “
Passive acoustic localisation of undersea gas seeps using beamforming
,”
Int. J. Greenhouse Gas Control
108
,
103316
(
2021
).
19.
B. J. P.
Bergès
,
T. G.
Leighton
, and
P. R.
White
, “
Passive acoustic quantification of gas fluxes during controlled gas release experiments
,”
Int. J. Greenhouse Gas Control
38
,
64
79
(
2015
).
20.
A.
Vazquez
,
R.
Manasseh
, and
R.
Chicharro
, “
Can acoustic emissions be used to size bubbles seeping from a sediment bed?
,”
Chem. Eng. Sci.
131
,
187
196
(
2015
).
21.
T. G.
Leighton
and
P. R.
White
, “
Quantification of undersea gas leaks from carbon capture and storage facilities, from pipelines and from methane seeps, by their acoustic emissions
,”
Proc. R. Soc. A
468
(
2138
),
485
510
(
2012
).
22.
M. S.
Longuet-Higgins
,
B. R.
Kerman
, and
K.
Lunde
, “
The release of air bubbles from an underwater nozzle
,”
J. Fluid Mech.
230
,
365
390
(
1991
).
23.
H.
Czerski
and
G. B.
Deane
, “
Contributions to the acoustic excitation of bubbles released from a nozzle
,”
J. Acoust. Soc. Am.
128
(
5
),
2625
2634
(
2010
).
24.
T. G.
Leighton
and
A. J.
Walton
, “
An experimental study of the sound emitted from gas bubbles in a liquid
,”
Eur. J. Phys.
8
(
2
),
98
104
(
1987
).
25.
H. C.
Pumphrey
and
J. E.
Ffowcs Williams
, “
Bubbles as sources of ambient noise
,”
IEEE J. Ocean. Eng.
15
(
4
),
268
274
(
1990
).
26.
M.
Minnaert
, “
XVI. On musical air-bubbles and the sounds of running water
,”
London, Edinburgh, Dublin Philosoph. Mag. J. Sci.
16
(
104
),
235
248
(
1933
).
27.
T. G.
Leighton
, “
The freely-oscillating bubble
,” in
The Acoustic Bubble
(
Academic
,
New York
,
1994
), pp.
129
286
.
28.
T. G.
Leighton
,
P. R.
White
, and
M. F.
Schneider
, “
The detection and dimension of bubble entrainment and comminution
,”
J. Acoust. Soc. Am.
103
(
4
),
1825
1835
(
1998
).
29.
M. R.
Loewen
and
W. K.
Melville
, “
An experimental investigation of the collective oscillations of bubble plumes entrained by breaking waves
,”
J. Acoust. Soc. Am.
95
(
3
),
1329
1343
(
1994
).
30.
J.
Li
,
P. R.
White
,
J. M.
Bull
,
T. G.
Leighton
,
B.
Roche
, and
J. W.
Davis
, “
Passive acoustic monitoring of a natural CO2 seep site—Implications for carbon capture and storage
,”
Int. J. Greenhouse Gas Control
93
,
102899
(
2020
).
31.
T. G.
Leighton
, “
The Rayleigh–Plesset equation in terms of volume with explicit shear losses
,”
Ultrasonics
48
(
2
),
85
90
(
2008
).
32.
T. G.
Leighton
,
The Acoustic Bubble
(
Academic
,
New York
,
1994
).
33.
H.
Medwin
and
M. M.
Beaky
, “
Bubble sources of the Knudsen sea noise spectra
,”
J. Acoust. Soc. Am.
86
(
3
),
1124
1130
(
1989
).
34.
G. B.
Deane
and
M. D.
Stokes
, “
The acoustic excitation of air bubbles fragmenting in sheared flow
,”
J. Acoust. Soc. Am.
124
(
6
),
3450
3463
(
2008
).
35.
T.
Monecke
,
S.
Peterson
,
M. D.
Hannington
,
M.
Anzidei
,
A.
Esposito
,
G.
Giordano
,
D.
Garbe-Schönberg
,
N.
Augustin
,
B.
Melchert
, and
M.
Hocking
, “
Explosion craters associated with shallow submarine gas venting off Panarea island, Italy
,”
Bull. Volcanol.
74
(
9
),
1937
1944
(
2012
).
36.
S.
Caliro
,
A.
Caracausi
,
G.
Chiodini
,
M.
Ditta
,
F.
Italiano
,
M.
Longo
,
C.
Minopoli
,
P. M.
Nuccio
,
A.
Paonita
, and
A.
Rizzo
, “
Evidence of a recent input of magmatic gases into the quiescent volcanic edifice of Panarea, Aeolian Islands, Italy
,”
Geophys. Res. Lett.
31
(
7
),
L07619
, (
2004
).
37.
S.
Inguaggiato
,
I. S.
Diliberto
,
C.
Federico
,
A.
Paonita
, and
F.
Vita
, “
Review of the evolution of geochemical monitoring, networks and methodologies applied to the volcanoes of the Aeolian Arc (Italy)
,”
Earth-Sci. Rev.
176
,
241
276
(
2018
).
38.
E.
Bagnato
,
E.
Oliveri
,
A.
Acquavita
,
S.
Covelli
,
E.
Petranich
,
M.
Barra
,
F.
Italiano
,
F.
Parello
, and
M.
Sprovieri
, “
Hydrochemical mercury distribution and air-sea exchange over the submarine hydrothermal vents off-shore Panarea Island (Aeolian arc, Tyrrhenian Sea)
,”
Mar. Chem.
194
,
63
78
(
2017
).
39.
M.
Molari
,
K.
Guilini
,
L.
Lins
,
A.
Ramette
, and
A.
Vanreusel
, “
CO2 leakage can cause loss of benthic biodiversity in submarine sands
,”
Mar. Environ. Res.
144
,
213
229
(
2019
).
40.
S.
Quale
and
V.
Rohling
, “
The European Carbon dioxide Capture and Storage Laboratory Infrastructure (ECCSEL)
,”
Green Energy Environ.
1
(
3
),
180
194
(
2016
).
41.
G.
Caramanna
,
N.
Voltattorni
, and
M. M.
Maroto-Valer
, “
Is Panarea Island (Italy) a valid and cost-effective natural laboratory for the development of detection and monitoring techniques for submarine CO2 seepage?
,”
Greenhouse Gases: Sci. Technol.
1
(
3
),
200
210
(
2011
).
42.
K.
Tsuchiya
,
K.
Ohsaki
, and
K.
Taguchi
, “
Large and small bubble interaction patterns in a bubble column
,”
Int. J. Multiphase Flow
22
(
1
),
121
132
(
1996
).
43.
C.
Martínez-Bazán
,
J. L.
Montañés
, and
J. C.
Lasheras
, “
On the breakup of an air bubble injected into a fully developed turbulent flow. Part 1. Breakup frequency
,”
J. Fluid Mech.
401
,
157
182
(
1999
).