Next Article in Journal
Tracking of Fin Whales Using a Power Detector, Source Wavelet Extraction, and Cross-Correlation on Recordings Close to Triplets of Hydrophones
Previous Article in Journal
Adaptive Convolution Kernels Construction Based on Unsupervised Learning for Underwater Acoustic Detection
Previous Article in Special Issue
Triaxial Experimental Study of Natural Gas Hydrate Sediment Fracturing and Its Initiation Mechanisms: A Simulation Using Large-Scale Ice-Saturated Synthetic Cubic Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bottom-Simulating Reflectors (BSRs) in Gas Hydrate Systems: A Comprehensive Review

1
Beijing International Center for Gas Hydrate, School of Earth and Space Sciences, Peking University, Beijing 100871, China
2
School of Environment and Energy, Peking University Shenzhen Graduate School, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2025, 13(6), 1137; https://doi.org/10.3390/jmse13061137
Submission received: 13 May 2025 / Revised: 1 June 2025 / Accepted: 4 June 2025 / Published: 6 June 2025
(This article belongs to the Special Issue Advances in Marine Gas Hydrates)

Abstract

The bottom-simulating reflector (BSR) serves as an important seismic indicator for identifying gas hydrate-bearing sediments. This review synthesizes global BSR observations and demonstrates that spatial relationships among BSRs, free gas, and gas hydrates frequently deviate from one-to-one correspondence. Moreover, our analysis reveals that more than 35% of global BSRs occur shallower than the bases of gas hydrate stability zones, especially in deepwater regions, suggesting that the BSRs more accurately represent the interface between the gas hydrate occurrence zone and the underlying free gas zone. BSR morphology is influenced by geological settings, sediment properties, and seismic acquisition parameters. We find that ~70–80% of BSRs occur in fine-grained, grain-displacive sediments with hydrate lenses/nodules, while coarse-grained pore-filling sediments host <20%. BSR interpretation remains challenging due to limitations in traditional P-wave seismic profiles and conventional amplitude versus offset (AVO) analysis, which hinder accurate fluid identification. To address these gaps, future research should focus on frequency-dependent AVO inversion based on viscoelastic theory, multicomponent full-waveform inversion, improved anisotropy assessment, and quantitative links between rock microstructure and elastic properties. These innovations will shift BSR research from static feature mapping to dynamic process analysis, enhancing hydrate detection and our understanding of hydrate–environment interactions.

1. Introduction

The bottom-simulating reflector (BSR), as the most prominent seismic indicator of gas hydrate-bearing sediments (GHBS), serves three critical scientific and practical roles: ① providing the fundamental constraint (distribution and GHBS thickness) for estimating hydrate reservoir volumes in conjunction with well log and/or core data, ② delineating geohazards from dissociation-induced seafloor instability, and ③ resolving methane release scenarios for climate modeling. However, critical uncertainties remain regarding the formation mechanisms of the BSR, coupling with underlying free gas systems, and dynamic responses to environmental perturbations. Resolving these gaps through integrated geophysical and numerical approaches is essential for advancing hydrate system predictability.
BSR studies originated alongside early gas hydrate theories. Trofimuk et al. first proposed natural hydrate accumulation models [1], while seismic surveys (e.g., Blake Ridge) identified anomalous reflectors parallel to the seafloor but oblique to the strata [2,3]. However, due to the limited resolution of single-channel P-wave data and the lack of in situ drilling validation, the formation mechanism of BSR was not fully understood.
Multichannel seismic acquisition and advanced inversion methods (e.g., travel time [4,5] and full waveform inversion [6,7,8,9]) revealed a diagnostic velocity structure: a high-velocity hydrate-bearing layer overlying a low-velocity free gas zone [10]. AVO analysis revealed BSRs’ positive amplitude-versus-offset trends, though these responses were later shown to depend nonlinearly on hydrate/gas saturations. Critically, ODP Legs 164/204 drilling refuted the Hydrate Wedge Model [10], demonstrating that impedance contrast from high-saturation hydrates alone cannot generate strong reflections. Instead, BSR was confirmed to result from the presence of free gas beneath the base of the gas hydrate stability zone (BGHSZ), where even ~5% gas saturation reduces P-wave velocity sharply [11], and amplifies reflectivity [12]. These insights established the Free Gas Zone Model [13] as the dominant paradigm.
Since the early 21st century, BSR research has entered a more refined and multidisciplinary stage, characterized by three major trends:
First, the use of 3D seismic data significantly enhanced stratigraphic resolution [14]. When combined with multi-wave and multi-component techniques and AVO attribute analysis, researchers developed BSR identification frameworks based on elastic parameters such as P- to S-wave velocity ratios and Poisson’s ratio [15,16]. Further studies have shown that the microstructural occurrence of hydrates in sediments (e.g., cementation, pore-filling, or grain-contact modes) directly affects elastic responses: hydrates acting as cement or in grain contact enhance both P- and S-wave velocities, whereas pore-suspended hydrates predominantly increase only P-wave velocity [17,18]. This distinction provides a theoretical basis for inferring hydrate occurrence mechanisms.
Second, research has expanded from static property analysis to dynamic process monitoring. Methane plume imaging has confirmed that gas released from hydrate dissociation migrates vertically along faults, forming high-amplitude reflective zones [19]. The spatial distribution of these features is closely associated with the boundaries of the hydrate stability zone [20], offering indirect evidence for BSR identification.
Third, studies on viscoelastic absorption characteristics have revealed marked regional variability. In some areas, hydrate-bearing sediments exhibit strong seismic attenuation [21], while in others, weak attenuation is observed [22,23,24]. These contrasting results may stem from complex interactions among hydrate occurrence modes, free gas distribution, and local geological conditions.
Despite significant advances in BSR research, four critical knowledge gaps persist: ① the limited generalizability of localized case studies, ② persistent ambiguities in BSR identification accuracy, ③ poorly constrained spatial correlations between BSRs and BGHSZs, and ④ unresolved formation mechanisms for double or multiple BSRs.
To systematically address these challenges, this review undertakes a comprehensive approach that includes: ① synthesizing global BSR occurrences across active and passive margins, ② diagnosing methodological and interpretive anomalies in existing datasets, ③ evaluating the relative controls of thermal, compositional, and tectonic factors on BSR morphology, and ④ quantifying global BSR–BGHSZ relationships through integrated seismological and thermodynamic modeling. By bridging theoretical predictions with observational inconsistencies, these efforts aim to standardize BSR interpretation frameworks, thereby refining hydrate resource estimates and improving constraints on climate-sensitive methane release dynamics.

2. Seismic Identification and Global Distribution of BSRs

BSR is a distinctive seismic reflection with negative polarity, typically subparallel to the seafloor and often crosscutting underlying stratigraphy. It generally forms under conduction-dominated thermal regimes rather than convective environments [25]. The strength of the reflection can be quantified using the reflection coefficient (R) or amplitude (A):
R = Z 2 Z 1 Z 2 + Z 1 = A 2 A 1
Z = ρ v
where Z represents acoustic impedance, calculated as the product of density (ρ) and P-wave velocity (v), and subscripts 1 and 2 correspond to the incident and reflected layers, respectively. The absolute value of R is positively correlated with reflection amplitude and typically falls within the range of −1 to 1 [26].
BSR arises from the acoustic impedance contrast between hydrate-bearing sediments and the underlying free gas zone. When gas hydrate saturation exceeds ~40%, it significantly increases the P-wave velocity in the overlying sediments (typically 1.8–2.5 km/s) due to grain-hydrate interaction [27]. Meanwhile, the presence of free gas in the underlying layer can significantly decrease the P-wave velocity (as low as 1.2–1.5 km/s) [11]. This sharp velocity contrast makes the BSR one of the most reliable seismic indicators for detecting GHBS [25]. A particular type of BSR, referred to as double or multiple BSRs, displays vertically stacked reflection patterns. These features are interpreted as the result of hydrate system re-stabilization following thermodynamic disequilibrium events and complex gas composition [28,29,30]. The global distribution of such multi-BSRs offers critical insights into the dynamic response of gas hydrate systems to climatic and geologic forcing over time.

2.1. Seismic Identification of BSR

Accurate seismic identification of BSR is critical for gas hydrate research, as misleading or pseudo-reflections may complicate BSR interpretation. Figure 1 presents representative seismic profiles from major global gas hydrate regions, emphasizing structural and stratigraphic features indicative of hydrate presence.

2.1.1. Seismic Characteristics of BSRs

BSR formation primarily depends on the regional thermal regime. In conduction-dominated systems (the most common scenario), the BSR parallels the seafloor or isotherms. In convection-dominated systems (e.g., fluid-venting zones), the BSR often deviates from seafloor geometry (see Section 4.1.3 described later). The BSR represents a phase boundary (gas hydrate stability zone, GHSZ) rather than a lithological contact, allowing it to cross-cut dipping strata. In terms of polarity, the BSR is characterized by a high-impedance hydrate-bearing layer overlaying a low-impedance free gas zone, resulting in a negative polarity reflection—an inverse polarity signature that contrasts with diagenetic BSRs, which typically exhibit positive polarity [31,32].

2.1.2. Seismic Anomalies Associated with BSR

The identification of the BSR on seismic profiles should rely on a combination of multiple seismic anomalies rather than a single seismic attribute. Beyond the characteristic negative-polarity BSR, several associated reflections commonly occur: ① a strong positive-polarity reflector beneath the BSR, marking the impedance contrast at the base of the free gas zone [33]; ② a positive-polarity reflector above the BSR in some cases, resulting from the hydrate-bearing layer’s upper boundary; and ③ acoustic blanking zones—both below the BSR (due to gas attenuation in chimneys or diapirs) and above it (caused by amplitude suppression in homogeneous hydrate-bearing sediments) (Figure 1). These reflection events typically parallel the seafloor [34].

2.1.3. Seismic Attribute Analysis of BSR

Conventional seismic profiles derived from raw data can effectively illustrate amplitude variations and stratigraphic geometry. However, they provide limited information on physical properties, lithology, and fluid content. As a result, distinguishing fluid differences in similar lithologies becomes difficult, which may lead to misinterpretation of BSRs in some cases. In contrast, seismic attribute analysis enables more detailed and accurate characterization of subsurface features by extracting amplitude, frequency, phase, and other derived parameters [35].
Seismic attributes can be derived from either post-stack or pre-stack seismic data. Post-stack attributes primarily include instantaneous amplitude, instantaneous phase, and instantaneous frequency [36], while pre-stack attributes are mainly extracted from Amplitude Versus Offset (AVO) analysis.
Figure 1. Seismic profiles of representative global gas hydrate regions: (a) Blake Ridge, modified from Hornbach et al. [37]; (b) Nyegga, Norway, modified from Hjelstuen et al. [38]; (c) Barents Sea, modified from Vadakkepuliyambatta et al. [39]; (d) Shenhu Area, South China Sea, modified from Jin et al. [40]; (e) Gulf of Mexico, modified from Portnov et al. [41]; and (f) Makran Accretionary, modified from Liu et al. [42].
Figure 1. Seismic profiles of representative global gas hydrate regions: (a) Blake Ridge, modified from Hornbach et al. [37]; (b) Nyegga, Norway, modified from Hjelstuen et al. [38]; (c) Barents Sea, modified from Vadakkepuliyambatta et al. [39]; (d) Shenhu Area, South China Sea, modified from Jin et al. [40]; (e) Gulf of Mexico, modified from Portnov et al. [41]; and (f) Makran Accretionary, modified from Liu et al. [42].
Jmse 13 01137 g001
Instantaneous amplitude, or reflection strength, represents the energy of a seismic signal at a given time and is always positive and phase-independent. The BSR at the interface between hydrate-bearing layers and underlying free gas zones typically exhibits strong amplitudes, while hydrate-bearing layers themselves often produce weak reflections due to their acoustic homogeneity. Compared with conventional profiles, amplitude attributes more clearly highlight high-contrast reflections linked to the BSR and hydrate accumulations [43]. The instantaneous phase reflects time-dependent wave phase changes and helps track reflection continuity. Phase shifts occur as seismic waves pass through lithological boundaries, making the BSR more distinguishable in phase sections. Since phase attributes are independent of amplitude, they are useful for identifying weak or discontinuous BSR, although their visibility decreases when the BSR is nearly parallel to bedding planes [44,45]. Instantaneous frequency, the time derivative of the instantaneous phase, detects low-frequency shadows beneath gas-saturated sediments, complementing amplitude data for gas/hydrate differentiation [36].
AVO analysis identifies lithology and potential hydrocarbon zones by investigating changes in reflection amplitude with offset (or incidence angle). In GHBS, BSRs frequently exhibit positive AVO anomalies, where reflection amplitude strengthens with increasing offset. However, this response depends strongly on hydrate/gas saturation and thickness, making AVO responses alone insufficient for accurately describing elastic properties [13,46,47,48,49]. To address this limitation, AVO attribute analysis focuses on hydrocarbon indicators such as the intercept and fluid factor. The intercept, representing the zero-offset P-wave reflection coefficient, is useful for identifying BSRs and inferring their geological nature. The fluid factor, highly sensitive to free gas content, plays a key role in distinguishing between hydrate-bearing sediments and underlying gas zones [50].
In practice, reliable BSR interpretation demands an integrated approach, combining multiple seismic attributes with drilling and well-logging data (see Table 1 for a comparative summary).

2.2. Global Distribution of BSR

Based on a systematic statistical analysis, this study delineates the global distribution characteristics of BSR associated with GHBS. Figure 2 provides an intuitive visualization of the spatial locations, distribution patterns, and morphological features of all identified BSR sites. To support this analysis, we systematically compile key parameters—including Tectonic Geological Features (TGF), pore habit, sediment features, seismic morphology types of BSRs (SMT), geological setting types of BSRs (GST), BSR depth (ZBSR, mbsf), water depth (WD, m), seafloor temperature (Tsf, °C), and geothermal gradient (GG, °C/km)—organized by region in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8 (Atlantic: Table 2; Pacific: Table 3; Arctic: Table 4; Antarctic: Table 5; Indian Ocean: Table 6; inland seas/lakes: Table 7; continental permafrost: Table 8). Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8 serve as the data backbone of this research, with their core value reflected in two key aspects: ① providing a comprehensive data foundation for the analytical calculations and discussions in subsequent sections and ② establishing a standardized reference database to facilitate further research within the academic community. A one-to-one correspondence between the codes in Figure 2 and those in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8 confers explicit spatial traceability on the dataset, markedly increasing its practical value and multidimensional analytical potential of the dataset. For the sediment types associated with BSR, it is assumed that the representative lithology at the BSR location corresponds to the stratigraphy within 50 m above it. In the absence of specific sedimentological data for certain locations, lithological information from nearby sites can be obtained from reports of the Ocean Drilling Program (ODP), Deep Sea Drilling Project (DSDP), or Integrated Ocean Drilling Program (IODP). These reports are accessible through the International Ocean Drilling Data Archive. While the lack of co-located data introduces some uncertainty, this approach provides the best available estimates for sedimentological characteristics. Through this statistical evaluation, several noteworthy phenomena have been identified, warranting further discussion. These observations can be categorized into four main types, as outlined in the following subsections.

2.2.1. Depth Distribution of BSR

The depth distribution of BSR is distinctly right-skewed (skewness = +1.49), with approximately 68% of observations occurring between 200 and 500 mbsf, and a mean depth of ~365 mbsf. Notably, BSRs deeper than 600 mbsf are predominantly found in rapidly subsiding basins with low geothermal gradients (typically <30 °C/km), suggesting that the BSR burial depth is primarily controlled by the interplay between geothermal gradient and water depth. However, the uppermost sediments (<100 mbsf) often lack the conditions necessary for BSR formation due to the significant consumption of methane by anaerobic oxidation of methane (AOM) processes mediated by microbial communities [52]. Nonetheless, field observations identify two exceptional cases where shallow BSRs develop: ① seepage-induced BSR shoaling and ② landward pinch-out of shallow BSRs.
The first type, BSR shoaling in seepage environments, is characterized by significant upward displacement of the BSR beneath mud volcanoes or pockmarks, occasionally intersecting the seafloor. This phenomenon has been reported in regions such as the El Arraiche Mud Volcano Field [53], Congo-Angola Margin [54], Husmus hydrocarbon field in Mid-Norway [55], and Rio Grande Cone on the Brazilian continental margin [56]. The primary driver of this process is the sustained migration of deep-sourced thermogenic gas, facilitated by mud diapirism and gas chimney systems, which transport methane to shallower depths. Additionally, localized high-intensity hydrothermal activity alters pressure–temperature (P-T) conditions, further contributing to BSR shoaling.
The second type, landward pinching-out of the shallow BSR, is primarily observed at the landward edges of hydrate-bearing regions, such as the Hydrate Ridge [57] and offshore Mauritania [58]. This phenomenon is controlled by multiple geological factors. First, a decrease in water depth toward the continental shelf modifies local P-T conditions, influencing hydrate stability. Second, high methane flux combined with highly permeable turbidite layers facilitates methane accumulation in shallow sediments, thereby elevating the sulfate–methane transition zone (SMTZ). Third, localized thermal perturbations induced by the high thermal conductivity of salt diapirs increase the geothermal gradient, further affecting BSR stability. A particularly important concern associated with BSR outcrops, where the BSR intersects the seafloor, is their positioning at the feather edge of the hydrate stability zone. This makes them highly susceptible to seafloor temperature fluctuations, increasing the risk of warming-induced hydrate dissociation. As a result, such settings are more prone to methane release, posing significant geohazard risks and environmental implications.

2.2.2. Temperature Distribution of BSR

The temperature at the BSR ranges from 10.7 °C to 27.8 °C, with a mean value of 17.2 °C. Fifty percent of the data (interquartile range) fall between 14.8 and 19.7 °C, and the distribution shows a moderate positive skewness (skewness = +0.66). Notably, several high-temperature outliers (>25 °C) are identified, primarily concentrated in tectonically active regions such as the Lima Basin [59] and the offshore area of the western Indian Ocean near Tanzania [60]. What is more, in regions such as the Western Marmara Sea, Western Indian Ocean offshore Tanzania, and Site 860 in Chile, BSRs have been observed at depths significantly deeper than the predicted BGHSZ for pure methane systems. Thermodynamic analyses suggest these anomalies correspond to structure II (sII) gas hydrates, characterized by enhanced thermal stability due to their distinct clathrate cage architecture, enabling persistence under elevated temperature regimes.

2.2.3. Geothermal Gradient Distribution for BSR Occurrence

Statistical analysis indicates that the geothermal gradient in global BSR-bearing regions ranges from 19 to 130 °C/km, with a mean value of 46.8 °C/km and a moderately right-skewed distribution (skewness = +1.86). The interquartile range is 32–52 °C/km, consistent with the global average thermal conduction regime of marine sediments. Notably, abnormally high geothermal gradients (>80 °C/km) are concentrated in areas such as the El Arraiche Mud Volcano Field [61] and the Scotia Sea [62], which are associated with intense heat flow upwelling within volcanic rift systems or nascent faulted margins. In these environments, elevated heat flux significantly raises the BGHSZ, resulting in a shallower BSR depth (mbsf). This effect underscores the dominant influence of tectonic activity on geothermal regimes.
Statistical comparisons further demonstrate that TGF is a primary control on the spatial variability of geothermal gradients: active continental margins—typically associated with subduction, rifting, or volcanism—exhibit an average geothermal gradient of approximately 55 °C/km, while passive continental margins—characterized by tectonic stability—have significantly lower values, averaging around 40 °C/km. Particularly low gradients (<32 °C/km) are observed in thickly sedimented basins such as the Blake Ridge, where massive sediment accumulations act as a thermal blanket, effectively impeding the upward transfer of deep heat [63,64,65,66,67,68].
The strong coupling between the geothermal gradient and TGF is further reflected in its negative correlation with BSR depth (Pearson’s r ≈ –0.48), indicating that areas with high heat flow tend to have shallow BSRs. These observations reveal a global bimodal distribution pattern of BSRs:
  • “Shallow BSRs with high geothermal gradients”, typically occurring in venting settings (e.g., gas chimneys or mud diapirs), where rapid fluid migration thins the gas hydrate stability zone and increases subsurface temperatures.
  • “Deep BSRs with low geothermal gradients”, commonly found in thick sedimentary basins, where a stable thermal regime allows the BGHSZ to extend below 300–600 mbsf.

2.2.4. Seismic Morphologies of BSR

A global statistical analysis of BSR seismic morphologies reveals four dominant types with distinct distribution patterns:
Continuous BSRs account for 37.50% of the total in the statistics and are mainly distributed along passive continental margins. These BSRs typically appear as continuous reflections in seismic profiles, parallel to the seafloor, and intersect with the stratigraphic reflections. A typical example is found in the Blake Ridge [25,51,69,70].
Discontinuous BSRs are the most widespread globally, comprising 54.69% of the total. They are primarily located along active continental margins and manifest as segmented, discontinuous lateral reflections, following the topography of the seafloor [51,69].
Clustered BSRs are rare, making up only 6.25% of the total. They are primarily localized at the crests of domes or anticlines. In such structural settings, the lateral migration of free gas is impeded, promoting its vertical accumulation beneath the structural crest. This process results in the entrapment of gas within discrete stratigraphic layers, producing a sequence of vertically stacked, high-amplitude reflections beneath a BSR that is roughly parallel to the seafloor. They often occur in areas of folding or salt-related domes, indicating a high concentration of natural gas hydrates and free gas beneath [41,71,72,73,74,75,76].
Pluming BSRs are also infrequent, accounting for just 1.56% of the total, such as in the Gulf of Mexico GC-955 [77]. These BSRs represent a unique type of continuous seismic reflection, typically found in localized areas associated with surface mud volcanoes or salt domes. They are smaller in scale and generally occur at relatively shallow depths, typically less than 1200 mbsf [51]. In seismic profiles, their distribution is not parallel to the seafloor, and their shape appears semicircular. They are often accompanied by acoustic blank zones [77].
Table 2. Global statistical analysis of known BSR data (Atlantic).
Table 2. Global statistical analysis of known BSR data (Atlantic).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
1/
2
Continental margin of Brazil (Foz do Amazonas Basin/Rio Grande Cone)PCM(Channel-levee turbidite systems and MTDs; 85% silt, 13% clay, 0.15% sand [78])/(fine-grained siliciclastic material [56])NADisc./
Cont. [56]
Stru./
Vent. [56]
450 [56]/
(200–300) [56])
(600–2800 [56])/
(500–3500 [56])
NA30–34 [79]NANA
3/
4/
5/
6
Gulf of Mexico (Green Canyon/Perdido Canyon/Terrebonne/Orca Basin)ACM20–50% silt, 50–80% clay, <1% sand [80]Nodules [81]Clus./
Clus./
Disc./
Plum. [51,77]
Stru./
Stru./
Stru./
Vent. [51,77]
450 [77]/
NA/
NA/
NA
2000 [77]/
NA/
NA/
NA
NANANANA
7/
8/
9
Blake Ridge (Site 533/Site 995/Site 997)PCM(11–28% silt, 70–86% clay, 0.3–3% sand)/(Nannofossil-rich clay and claystone, with sediments exhibiting moderate to intense bioturbation)/(Homogeneous dark green clay and claystone, rich in diatoms) [63,64,65,66,67,68]Nodules [66]Cont. [64]/
(Cont.-Plum.) [68]/
Cont. [67]
Stra. [64]/
Stra. [68]/
Stra. [67]
600 [64]/
450 [68]/
464 [67]
3184 [64]/
2779.5 [68]/
2770.1 [67]
2.15 [64]/
3.6 [68]/
3.39 [67]
36 [64]/
33.5 [68]/
36.8 [67]
584.34/
553.22/
505.41
627.46/
600/
547.7
10Carolina Trough (Site 991)PCM15% silt, 85% clay, 0% sand [82]Nodules [83]Disc. [84]Stru. [84]300 [84]1500 [84]3.1 [82]NANANA
11Continental Rise (Site 107)PCM25% silt, 75% clay, 0% sand [85]NACont. [86]Stra. [86]400 [86]2700–3400 [86]NA36–51 [86]NANA
12Namibe Basin (Site 1080)PCMDiatom bearing 5–10% silt, 30–60% clay, 0% sand [87]Nodules [88]Diag. [87]NA250 [87]1000 [87]4.4–4.6 [88]NANANA
13Punta del Este BasinPCMNANACont. [89]Vent. [89]400 [89]350–2200 [89]NANANANA
14/
15
Norwegian Continental Slope (Husmus/Nyegga)PCM(Soft clay) [55]/
(45% silt, 54% clay, 1% sand) [90]
Nodules [91]Disc./
Cont. [55]
Vent./Vent. [55]5/
280 [55]
330/
730 [55]
(6–7)/
(−0.7) [55]
NA/
50 [92]
NANA
16/
17/
18/
19/
20
Newfoundland (Makkovik Slope/Hamilton Spur/Haddock Channel/Mohican Channel/Barringto)PCMClosely related to the drift deposition area [93]NADisc./
Cont./
Cont./
Cont./
Disc. [93]
Stru./
Stra./
Stra./
Stra./
Stra. [93]
443/
335/
375/
290/
378 [93]
(620–2555)/
(1075–2100)/
(1700–2150)/
(1324–2875)/
(2220–2850) [93]
(2.1–4.15) [93]32 [93]NANA
21Newfoundland Sackville SpurPCMClosely related to the drift deposition area [93]NADisc. & (Doub/Mult.) [94]Stra. [94]350 [94]1125 [94]3.9 [95]35.8 [95]314.94360.82
22Offshore MauritaniaPCMFine-grained turbidites and nannofossil-rich muds [58]NADisc. [58]Stru. [58]83 [58]711 [58]8 [96]32 [96]39.96105.52
23Colombia Basin (Site 502)ACMMostly nannofossils and clay; 25–75% clay [97,98]NACont. [99]Stru. [99]485 [99]2640 [99]6 [76]20 [76]823.06905.67
24PanamaACM28% silt, 70% clay, 2% sand [100]NACont. [101]Stru. [101]180–300 [101]1800–2800 [101]NA59 [101]NANA
25El Arraiche Mud Volcano FieldPCMMud breccia & clay [61]Nodules [102]Cont. [53]Vent. [53]0 [53]380 [53]10 [53]110 [53]NANA
26Congo Deep Sea Fan (Site 1076)PCM25–75% silt, 25–75% clay, 0% sand [54,103]Nodules [54]Disc. [104]Stru. [104]220 [104,105]2980 [104,105]2.5 [105]80 [105]241.87260.57
27Niger Delta FrontPCM80% fines, 20% sand [106]Nodules/
lenses [107]
Disc. [108]Stra. [108]300–380 [108]2400–2900 [107,108]4.45–4.53 [107]72 [107]NANA
28Offshore UK (Faroe-Sheland Basin)PCMCoarse-grained sediments, linked to paleochannelsNACont. [109] Stru. [109]300–350 [109]>750 [109]NA36.5 [109]390 [109]NA
Note: NA = Not Available; PCM = Passive Continental Margin; TGF = Tectonic Geological Features; ACM = Active Continental Margin; MCM = Mixed Continental Margin; TRL = Tectonic Rift Lake; TPB = Transition Plate Boundary; ISB = Inland Sedimentary Basin; CC = Continental Collision; SMT = Seismic Morphology Types of BSRs; Cont. = Continuous BSR; Disc. = Discontinuous BSR; Clus. = Clustered BSR; Plum. = Pluming BSR; Diag. = Diagenetic BSR; GST= Geological Setting Types of BSRs; Stru. = Structural Setting; Stra. = Stratigraphic Setting; Vent. = Venting Setting; ZBSR = Depth of BSR; WD = Water Depth; Tsf = Seafloor Temperature; GG = Geothermal Gradient; ZBGHSZ = Depth of BGHSZ in Salt Water (3.5 wt.% NaCl); ZBGHSZ* = Depth of BGHSZ in Pure Water. The notes for Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8 are identical to those for Table 2.
Table 3. Global statistical analysis of known BSR data (Pacific).
Table 3. Global statistical analysis of known BSR data (Pacific).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
29Acapulco (Site 491)ACM50% silt, 30% clay, 20% sand [110]Nodules [110]Cont. [110]Stru. [110]380 [101]2000–3800 [101]NA21 [101]NANA
30Hydrate Ridge (Site 1245)ACM50% silt, 50% clay [111]Layered [111]Cont. [57]Stru. [57]134 [57]870 [57]4 [57]54 [112]150.91180.32
31Cascadia Margin (Site 889)ACM30–50% silt, 40–70% clay, 0–4% sand [113,114]Nodules [115]Disc. [114]Stru. [114]224 [114]1313 [114]2.7 [114]54 [114]243.22271.66
32Lima Basin (Site 685)ACMMainly composed of diatom mud and calcareous mudstone; 45% silt, 55% clay [59,116]Nodules [59]Disc. [59]Stru. [59]612 [59]5070 [59]1.5 [59]43 [59]590.16625.49
33Nankai TroughACM26–36% silt, 1–3% clay, 61–72% sand [117,118]Pore-filling [119]Disc. [118]Stru. [118]204 [119]945 [119]3 [119]46 [120]225.55260.52
34Sea of Okhotsk (Site 796)MCM73–79% silt, 20.4–26.3% clay, 0–5% sand [121]Nodules [122]
/lenses [123]
Disc. [124]Stru. [124]450 [125]1300 [125,126]2.5 [126]35 [126]402.49448.72
35Ulleung BasinACM100% <75 μm, median grain size: 2.3–3.0 μm [127]Nodules/
Lenses [128]
Disc. [128]Stru.-Stra. [128]150 [128]2100 [127,128]NANANANA
36/
37/
38/
39
Shenhu Area (SH1/SH2/SH3/SH7)PCM(Clay & Silt)/
(70% silt, 25% clay, 5% sand)/
(Clay & Silt)/
(Silt & Sand) [129,130]
Pore-filling [130]; Nodules [131]Disc./
Cont./
Cont./
Cont. [130]
Stra./
Vent./
Vent./
Vent. [130]
219/
221/
204/
181 [130]
1500 [130]5.2/
4.84/
5.53/
6.44 [130]
47.53/
46.95/
49.34/
43.65 [130]
222/
229/
206/
184 [130]
276.4/
288.72/
257.78/
272.06
40Hikurangi Trough (Site U1517)ACM40% silt, 58% clay, 2% sand [132]Lenses [133](Cont.) & (Doub./Mult.) [134]Stra. [134]165 [134]725 [134]5.32 [135]39.8 [135]130.58175.57
41/
42/
43
Chile (Site 859/Site 860/Site 861)ACM(Silty clay & Clayey silt)/
(44% silt, 49% clay, 7% sand)/
(Coarse-grained clastic sediment) [136,137]
NADisc./
Cont./
Disc. [136]
Stru./
Stru./
Stru. [136]
100/
200/
250 [138]
2741.2/
2145.9/
1652 [136]
NA/
2.5 [137]/
NA
NA/
103.5 [137]/
NA
NA/
160.2/
NA
NA/
174.46/
NA
Table 4. Global statistical analysis of known BSR data (Arctic).
Table 4. Global statistical analysis of known BSR data (Arctic).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
44Sverdrup BasinPCMSandstone [139,140]NADisc. [141]Stru. [141]900 [141]NANA20–40 [139]NANA
45Fram Strait (Site 986)PCM30% silt, 65% clay, 5% sand [142]NADisc. [143]Stru. [143]200 [143]2500 [143]0 [143]75 [143]274.71294.68
46Aleutian Trench (Site 186)ACM25% silt, 45% clay, 5% spicules, 25% diatoms [144]NADisc. [145]Stru. [145]926 [145]4500 [145]1 [145]20 [145]1306.051385.6
47Shirshov Ridge/KoryakACM11.3–25.8% silt, 5.6–12.1% clay, 62.1–83.1% sand [146]NADisc. [147]Stru. [147]200–500 [147]1500–3000 [147]NA58 [147]NANA
48Barents Sea (Site 7316/03-U-01)PCM20% silt, 60% clay, 20% sand [148]NAClus. [39,149]Stru. [39,149]180 [150]345 [150]2 [149]NANANA
49Offshore NW Greenland (Baffin Bay)PCMNANADisc. [151]Stru. [151]200 [151]625–720 [151]10–11.6 [151]40–59 [151]NANA
Table 5. Global statistical analysis of known BSR data (Antarctic).
Table 5. Global statistical analysis of known BSR data (Antarctic).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
50Ross Sea (Site 273)PCM40% silt, 59% clay, 1% sand [152]NADisc.& (Doub./Mult.: BSR1-SI hydrate; BSR2-sII hydrate) [153]Stru. [153]500 [153]880 [153]−1.5 [153]36 [153]442.96489.1
51Scan Basin (BSR1/BSR2/BSR3)ACMNANA((Cont.)/(Diag.: Opal-A/CT)/(Diag.: Opal-CT/Quartz)) & (Mult.) [62]Vent. [62]/
NA/
NA
145/
429/
1360 [62]
2100 [62]−0.4 [62]/
NA/
NA
130/
(60–70)/
NA [62]
148.25159.51
52Weddell Sea (Site 695)PCMNANADiag. [154]NA690 [154]1300 [154]NA52 [154]NANA
Table 6. Global statistical analysis of known BSR data (Indian Ocean).
Table 6. Global statistical analysis of known BSR data (Indian Ocean).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
53Gulf of Oman (Site 222)MCM64–76% clay, 24–36% silt, 0% sand [155]NADisc. [156]Stru. [156]680 [156]3000 [156]2 [156]30 [156]701.02753.51
54/
55/
56/
57/
58/
59
Krishna-Godavari Basin (NGHP-01-01/NGHP-01-07/NGHP-01-09/NGHP-01-17/NGHP-01-21)PCM(Carbonate oozes)/
(Clay with silt or sand beds)/
(Clay or silt)/
(Clay or silt)/
(Clay or silt with volcanic ash beds)/
(50–70% clay, 30–50% silt) [157,158]
Nodules/
Lenses [159]
Diag./
Disc./
Disc./
Clus./
Clus./
Disc. [158]
NA/
Stra./
Stru./
Stru./
Stra./
Stra. [158]
220/
188/
290/
160/
608/
160 [158]
2663/
1285/
1935/
1038/
1344/
1049 [158]
2.4/
5.21/
5/
6.5/
5.6/
6 [158]
52/
51/
51/
45/
19/
45 [158]
NA/
200.6/
268.81/
156.82/
645.98/
172.2
NA/
230.98/
298.85/
192.61/
740.87/
207.9
60Sunda ArcACMInterbedded sandstone, argillaceous siltstone, dolomitic limestone, and volcanic deposits form alternating high-low permeability sequences [160]NADisc. [160]Stru. [160]150 [160]1500–2200 [160]NANANANA
61/
62
Western Indian Ocean Offshore Tanzania (BSR1 (Biogenic)/BSR2 (thermogenic))ACMNA/
Hemipelagic units, slope channel deposits & turbidite sands [60]
NADisc./
Disc. [60]
Stru./
Stru. [60]
350/
450 [60]
1800/
2362.5 [60]
9/
0.7 [60]
28.2/
54 [60]
335.09/
366
393.24/
294.2
63North Carnarvon BsinPCMCarbonate ooze [161] NADisc. [161]Stru. [161]NA> 1000 [161]NA54–63 [161]258 [161]NA
Table 7. Global statistical analysis of known BSR data (Inland seas and lakes).
Table 7. Global statistical analysis of known BSR data (Inland seas and lakes).
CodeLocationTGFSediment FeaturesPore HabitSMTGSTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
64Black SeaMCMParticle size distribution: >0.1 mm (not detected), 0.01–0.1 mm (10–30%), <0.01 mm (70–90%) [162]Nodules and lenses [163](Disc.) & (Doub./Mult.) [164]Stru. [164]320 [164]1330 [164]9 [165]24.5 [165]293.45364.95
65Baikal LakeTRL70% clay, 5% silt, 25% sand [166]Nodules and lenses [163]Disc. [163]Stru. [163]160 [163]1600 [163]3.5 [163]NANANA
66Anaximander Mud VolcanoesACM56–67% clay, 19–30% silt, 14% sand [167]NADiag. [167]NA40–200 [167]2025 [167]13.75 [167]26–38 [167]NANA
67Nile Deep-Sea FanPCMFine grained turbidite sediments and MTDs [168]NADisc. [168]Stru. [168]190 [168]2443 [168]12.5 [168]40 [168]191.9230.81
68Western Marmara SeaTPB50–70% clay, 30–40% silt, 0–1% sand [169]Nodules [170]Cont. [171]Vent. [171]25 [170,171]680 [170,171]14.5 [170,171]39.3 [170]−324.95−232.16
Table 8. Global statistical analysis of known BSR data (Continental permafrost).
Table 8. Global statistical analysis of known BSR data (Continental permafrost).
CodeLocationTGFSediment FeaturesPore HabitSMTZBSR (mbsf)WD
(m)
Tsf
(°C)
GG (°C/km)ZBGHSZ (mbsf)ZBGHSZ* (mbsf)
69Qilian Mountains TibetCCSandstones (medium to fine-grained), siltstones, mudstones, and oil-bearing shales [172]Fracture filling in rock [173,174]NA133–396 [173]0NANANANA
70West Siberia BasinISBParticle size distribution: 8% at 0.2 mm, 4% at 0.5 mm, 4% at 0.8 mm, and 84% exceeding 1 mm. [175]NANA800 [175]0NANANANA
71MallikISBFine/medium sand, mean grain size: 149.9–502.5 μm [117,176]Pore-filling [176]NA1000 [176]0NANANANA
72Lena-Tunguska BasinISBNANANA800–2000 [139]0NANANANA
73Mt ElbertCC56–61% <75 μm, mean grain size: 0.07–0.074 mm [177,178]Pore-filling [117,176]NA850 [177]0NANANANA

3. BSR as Indicators for the Presence of BGHSZ and Gas Hydrate/Free Gas

The GHSZ is primarily controlled by pressure (P) and temperature (T), with temperature variations exerting a significantly greater impact on hydrate stability than pressure fluctuations [179]. BSR has long been recognized as the indicator for identifying gas hydrate occurrences; however, the precise correspondence between BSR depth and the BGHSZ remains debated. Some studies suggest that BSR depth accurately represents the BGHSZ [26,180,181,182,183,184], whereas others argue that BSR depth typically occurs shallower than the BGHSZ, indicating a systematic offset [185].

3.1. Comparison of BSR Depth and Predicted Depth of BGHSZ

To assess this relationship, we used the Colorado School of Mines Hydrate (CSMHYD) program to compute methane hydrate phase equilibrium curves under both 3.5 wt.% salinity and pure water conditions [186]. By integrating measured data from 34 global hydrate-bearing sites with observed BSR, we derived the sub-seafloor depths of the BGHSZ (in meters below seafloor, mbsf) for both scenarios: ZBGHSZ, 3.5 wt.% salinity (seawater); and ZBGHSZ*, pure water (derived from the datasets in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8 and visualized in Figure 3c,d) [187]. We then conducted linear correlation analyses comparing these predicted depths with observed BSR depths ( Z B S R ). Notably, we introduced ΔZ as a quantitative measure of the depth deviation between BGHSZ and BSR (Equation (3)). Considering the vertical resolution of seismic data (10–20 m), a 20 m threshold criterion was established: BSR was considered coincident with BGHSZ when ΔZ ≤ 20 m, while BSR was determined to be shallower than BGHSZ when this threshold was exceeded.
Δ Z = M i n i m u m ( Z B G H S Z Z B S R ,   Z B G H S Z * Z B S R )
In contrast to theoretical expectations, our analysis identified three sites where BSR depths exceeded the predicted BGHSZ, all corresponding to Type II gas hydrate systems (confirmed from the literature review, as described in Figure 3a). These anomalies were consequently excluded from further analysis. As illustrated in Figure 3b, the surveyed depths of BSRs coincide with the theoretical base of the BGHSZ at 64.5% of the surveyed sites. However, 35.48% exhibited BSRs shallower than the BGHSZs, indicating that BSR should not be treated as a definitive proxy for the BGHSZ. Instead, it more reliably marks either the base of the hydrate occurrence zone (BGHOZ) or the top of the free gas zone (TFGZ).
This discrepancy arises from the potential existence of free gas between BGHOZ and BGHSZ under two specific conditions:
  • Capillary pressure effects in fine-grained sediments can inhibit hydrate formation, allowing methane to persist in gaseous form [188].
  • Local variations in P-T conditions triggered by seasonal fluctuations, elevated bottom-water temperatures, or tectonic uplift can induce partial dissociation of gas hydrates, temporarily releasing free gas and creating conditions of disequilibrium [28,29].
Theoretical models demonstrated that BGHSZ, BGHOZ, and TFGZ coincide spatially only when methane concentration exceeds solubility thresholds and flux surpasses critical rates [185]. Moreover, we found that when the BSR depth is less than approximately 2700 mbsl, its deviation from the BGHSZ is relatively small (average ΔZ = 11.5 m); whereas when the BSR depth exceeds ~2700 mbsl, it is generally much shallower than the BGHSZ (average ΔZ = 112.2 m). It is important to note that the interpretation of this depth-dependent deviation must be based on reliable data quality and accurate BSR identification. Excluding potential errors from literature data or misinterpretation of BSR, the substantial ΔZ observed in deeper waters may be attributed to reduced methane flux caused by a sharp decline in organic matter input, preventing the BSR from fully developing down to the theoretical base of the stability zone. Accordingly, BSRs located above 2700 mbsl can be considered approximate indicators of the BGHSZ, whereas, in deeper regions, systematic deviations between the BGHSZ and BGHOZ must be accounted for.
These findings improve the accuracy of GHSZ delineation for resource evaluation while enhancing methane leakage monitoring capabilities. Rigorous validation of BSR interpretations remains essential, especially in deep-water environments where data uncertainties amplify.

3.2. BSR as an Indicator of Gas Hydrate and Free Gas Presence

The relationship between BSRs, free gas, and gas hydrates exhibits complex variability, with three primary scenarios requiring differentiation.

3.2.1. Gas Hydrate Presence Without Observable BSR

Drilling results from the Gulf of Mexico and Blake Ridge demonstrate that gas hydrates can exist without seismic BSR expression when free gas is absent below the hydrate layer [181,189]. This eliminates the acoustic impedance contrast needed for BSR generation, as hydrate-bearing sediments and water-saturated strata exhibit similar seismic properties.
When the thickness of the free gas zone causes destructive interference between seismic reflections from its upper and lower boundaries, the BSR signal may weaken or disappear. Specifically, the seismic waves traveling between these interfaces generate phase differences. If this travel path difference is an odd multiple of the wavelength, the reflected waves interfere destructively, eliminating the observable BSR signal [190].

3.2.2. False “BSR” Without Gas Hydrates Presence

The BSR is caused by contrasts in acoustic impedance; however, similar strong reflections can also be generated by other factors even in the absence of gas hydrates. High-saturation free gas zones can produce negative-polarity reflections mimicking BSRs through abrupt P-wave velocity reduction, even in hydrate-free systems [191]. Diagenesis forms pseudo-BSR features through two distinct mechanisms. Silicate diagenesis (e.g., opal-A→CT→quartz and smectite→illite transitions) generates acoustic impedance contrasts via pore compression and dehydration, forming sub-seafloor-parallel interfaces controlled by geothermal gradients [31]. These interfaces exhibit positive polarity and occur at depths significantly below the hydrate stability zone, distinguishing them from the hydrate-related BSR [62,154,192]. Carbonate diagenesis, observed in the Xisha Trough, Namibe Basin Site 1080, and Krishna-Godavari Basin NGHP-01-01, produces high-amplitude continuous reflectors with negative polarity and geometry akin to the hydrate-related BSR [32,87,158]. Liang et al. developed a Carbonate-Hydrate Identification Template (CHIT) using a hybrid matrix velocity model, confirming that these carbonate-derived reflectors are independent of thermobaric conditions, fundamentally differing from hydrate phase equilibria [32].
Mass-transport deposits (MTDs), formed by gravity-driven submarine slope failures, often display complex internal architectures. Variations in density and seismic velocity between displaced sediment blocks can generate strong reflections and polarity reversals. While these reflections may exhibit polarity characteristics similar to the BSR, their irregular geometries and discontinuous reflection patterns reflect sediment dynamics governed by slope morphology and are not related to the hydrate stability zone [168,193].

4. Controls on BSR Morphology in Gas Hydrate Systems

The morphology of BSR is influenced by three main factors: geological settings, sediment types, and seismic data acquisition methods. The first two reflect natural geological conditions, whereas the third arises from instrumental and methodological constraints.

4.1. Influence of Geological Settings

The geological settings in which BSR occurs can be broadly categorized into structural settings, stratigraphic settings, and venting settings. Each setting is governed by distinct geological characteristics and formation mechanisms [184], as illustrated in Figure 4.

4.1.1. BSRs in Structural Settings

Structural settings account for 58.46% of all documented BSR occurrences, making them the predominant category (as detailed in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8). However, structural deformation—including faulting, mud diapirism, and salt tectonics—typically limits their lateral continuity, resulting in smaller-scale BSRs. The gas responsible for forming the BSR in these environments often originates from deep, focused hydrothermal fluids. Based on its distribution and the gas hydrate accumulation process, the BSR in tectonic settings can be further classified into three types, as illustrated in Figure 5 [195].
The ridge-type BSR develops at structural highs and appears as a convex-shaped reflection (Figure 5a). In this setting, free gas ascends preferentially along the ridge axis and accumulates beneath the GHSZ, generating a well-defined BSR. Prominent examples include gas hydrate systems in the Sea of Okhotsk [124,125].
The buried anticline-type BSR forms within anticlines that are overlain by younger sediments (Figure 5b). This BSR typically exhibits a flat-lying seismic reflector, parallel to the seafloor, where upward-migrating gas accumulates beneath the overlying seal. Such configurations have been documented in several sedimentary basins, including the southwestern margin of the Kumano Basin in the Nankai Trough [195], the eastern Beaufort outer slope [196], and the South Makassar Basin in Indonesia [197].
The accretionary prism-type BSR is mainly found within accretionary wedges (Figure 5c). Its distribution is less influenced by local geological structures or seafloor topography. Instead, it is primarily controlled by intense and diffuse upward fluid flow driven by tectonic thickening and sediment loading. Representative examples have been reported in the Gulf of Mexico and Acapulco [110], as well as offshore Mauritania [58].
Clustered and discontinuous forms of the BSR are more commonly observed in structural settings (Figure 4(a1–a2,b1–b2)). Among the worldwide BSR observations, clustered BSRs account for 10.53% and discontinuous BSRs for 73.68% of all BSR occurrences identified in structural settings, as summarized in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8. Notably, all instances of the clustered BSR are found exclusively in such settings.
This prevalence can be attributed to the role of faults and fracture systems, which serve as vertical pathways for deep-sourced fluids. These conduits facilitate the focused migration and accumulation of gas, resulting in zones of high hydrate saturation near the BGHSZ, and ultimately lead to the development of the localized, clustered BSR [198,199].
In addition, tectonic activity often generates numerous faults that enable the upward movement of hot fluids from depth, which can inhibit hydrate formation and cause interruptions in the continuity of the BSR [200,201]. Furthermore, processes such as salt tectonics and folding can lead to tilting of the underlying strata. This structural deformation alters porosity and permeability at different positions along the BGHSZ/BGHOZ, resulting in uneven distributions of gas and hydrates and ultimately giving rise to the discontinuous BSR [184].

4.1.2. BSRs in Stratigraphic Settings

BSRs in stratigraphic settings account for 24.62% of all identified BSRs, as indicated in Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8. These BSRs lack obvious fluid venting structures or seismically detectable faults. The methane responsible for hydrate formation in these settings is more likely derived from in situ microbial degradation of organic matter [202] or from deep thermogenic gas migrating through high-permeability layers or undetectable fractures [203]. In these settings, continuous BSRs (56.25%) dominate over discontinuous BSRs (43.75%) with their distribution strongly influenced by the stratigraphic geometry relative to the BGHSZ (Figure 3c,d).
Uniform porosity/permeability along the BGHSZ/BGHOZ promotes homogeneous fluid distribution, forming continuous BSRs (e.g., Puke Ridge in the Hikurangi Trough [204], Site 997 at Blake Ridge [67], and Site NGHP-01-17 in the Krishna-Godavari Basin [158]). Heterogeneous sediment permeability leads to compartmentalized gas/hydrate accumulations, generating discontinuous BSRs (e.g., Sackville Spur offshore Newfoundland [94]).

4.1.3. BSRs in Venting Settings

Venting-related BSRs represent 16.92% of observed occurrences and are linked to focused fluid expulsion through mud volcanoes, gas chimneys, or hydrothermal vents. In venting environments, continuous BSRs and pluming BSRs are the most commonly observed types, comprising 81.82% and 9.09% of venting-related BSRs in the statistical region, respectively (Figure 4e1–e2). On seismic profiles, they typically exhibit circular or semi-circular structures, often accompanied by acoustic blanking zones in the surrounding sediments [51,205].

4.2. Influence of Sediment Types

Triangular grain-size distribution analysis (Figure 6) indicates that BSR preferentially occurs in fine-grained sediments with high clay (40–80%) and silt (20–60%) content, particularly in sediment types such as clay, silty clay, clayey silt, and sand–silt–clay. In contrast, BSR is rare or nearly absent in coarser sediments where sand content exceeds 50%, such as sand and silty sand.
Continuous BSRs predominantly develop in fine-grained sediments (clay, sandy clay, and silty clay), characterized by small grain sizes and high clay content (40–80%). This preference may arise from the higher sealing capacity of fine-grained materials, which facilitates stable and continuous reflections. In contrast, discontinuous BSRs exhibit broader sedimentological affinity, occurring in mixed grain-size deposits (e.g., sandy clay, sand–silt–clay, silty clay) with medium-to-fine textures. Their irregular distribution likely results from heterogeneous geological controls, including permeability variations and localized fluid migration pathways. Clustered BSR is relatively rare and is exclusively found within structural settings. Previous studies suggest that its occurrence is often associated with coarse-grained sediments such as channel sands [76].
Sediment type not only influences the seismic morphology of the BSR but also determines the pore habit of gas hydrates. In fine-grained sediments, high capillary pressure promotes hydrate growth within small pores, replacing sediment grains and forming discrete hydrate lenses. This pore habit is referred to as “grain-displacive” hydrate formation. In contrast, in clean, coarse-grained sediments, hydrates experience lower capillary pressure within larger pores, allowing them to fill or invade pores without altering the sediment framework. This behavior is known as “pore-invasive” or “pore-filling” hydrate formation [203,206]. Pore-invasive hydrates are primarily observed in the Nankai region, Shenhu area, and Arctic permafrost zone, characterized by extensive distribution but lower saturation, making them less economically viable for extraction.

4.3. Influence of Seismic Data Acquisition Methods

The BSR morphology is shaped not only by geological settings and sediment types but also by seismic data acquisition parameters, including sampling intervals, frequency, and survey line orientation. Smaller sampling intervals and higher frequencies enhance seismic resolution, which comprises vertical and horizontal components. The vertical resolution (h) approximates one-quarter of the seismic wavelength, while the horizontal resolution (r) can be expressed by the radius of the Fresnel zone (Equations (4) and (5)). The continuity of BSR is predominantly influenced by horizontal resolution; higher seismic frequencies result in shallower propagation depths and smaller Fresnel zone radii, thus improving horizontal resolution.
h = v 4 f
r = v · z 2 f
where v is the seismic wave velocity, f is the dominant seismic frequency, and z is the target depth.
Studies by Hillman and Crutchley demonstrated that low-frequency seismic data can better capture extensive features, such as gas–water interfaces, thereby presenting BSR as a continuous reflection. However, these data often lack sufficient detail to elucidate complex gas migration pathways. Conversely, high-frequency data provide more detailed images of the specific distribution of hydrates and free gas near the GHSZ base, yet these data generally exhibit lower amplitudes due to greater attenuation [69,207]. Consequently, a BSR appearing continuous in low-frequency data might appear discontinuous in high-frequency data from a lateral perspective. Vertically, this also implies that a BSR represents a transition zone of a certain thickness, within which hydrate progressively transitions into free gas, rather than a simple binary interface between a hydrate-bearing layer and a free gas-bearing layer as traditionally hypothesized.
Furthermore, the orientation of seismic survey lines significantly affects the continuity of BSR, particularly when sediment layers are inclined. If the seismic line is oriented parallel to the direction of sediment tilt, it traverses multiple inclined layers intersected by the BSR. Variations in porosity and permeability along the same isotherm, resulting from sediment tilt, typically cause BSR to appear discontinuous. Conversely, if the seismic line is perpendicular to the direction of sediment tilt, the distribution of porosity and permeability within the same stratigraphic layer is relatively uniform, leading to continuous BSR reflections. However, in this case, the absence of information regarding sediment inclination may result in a BSR that does not accurately reflect the true distribution of hydrates and free gas [69].
As shown in Section 4.1 and Section 4.2, disregarding the influence of seismic acquisition methods, both geological settings and sediment types jointly affect the morphology of the BSR. Continuous BSRs prevail in stable seepage environments with homogeneous lithology (e.g., fine-grained seal layers overlying permeable pathways). These settings facilitate fluid/gas migration and accumulation, yielding consistent gas hydrate formation. Discontinuous BSRs dominate tectonically active regions with heterogeneous sediments (e.g., sandy clay and sand–silt–clay mixtures). Here, structural deformation (e.g., faulting) disrupts gas distribution, while lithological variability further amplifies BSR segmentation. This multiscale analysis clarifies the interplay between sedimentation and tectonics in BSR formation, providing critical insights for gas hydrate prospecting and reservoir characterization.

5. Formation Mechanisms of Double and Multiple BSRs

Although BSR is widely distributed along continental margins, double or multiple BSRs are relatively rare. Global observational data indicate that their presence primarily occurs along the continental margin of western Norway [208,209], the Nankai Trough offshore Japan [195], the Manila subduction zone [210], Hydrate Ridge offshore Oregon [30], the Ross Sea in Antarctica [153], the Danube deep-sea fan in the Black Sea [165], the Tumbes Basin offshore Peru [211], the Hikurangi margin offshore New Zealand [212], the South China Sea [213,214,215], Sackville Spur offshore Newfoundland [94], the Mauritanian continental margin in West Africa [216], and the Makran accretionary wedge along the northwestern coast of the Indian Ocean [217].
Double BSRs typically reflect disequilibrium states in gas hydrate systems, with the primary reflector (BSR1) corresponding to equilibrium conditions under current or recent P-T conditions, whereas the secondary reflector (BSR2) represents a relict paleo-BSR formed due to changes in P-T conditions leading to partial dissipation [218]. In specific instances, differences in hydrate type may also result in the formation of double BSRs [219]. Typically situated near ridge crests, double BSRs appear as two approximately parallel reflectors that mimic the seafloor topography and display polarity reversals. Notably, BSR2 often exhibits weaker amplitude and poorer continuity than BSR1 [211,220]. High acoustic impedance anomalies above and adjacent to both BSR1 and BSR2 indicate accumulations of natural gas hydrates, whereas low acoustic impedance anomalies directly below these reflectors suggest the presence of free gas accumulations [221]. Fundamentally, double/multiple BSRs reflect the dynamic response of gas hydrate systems to variations in P-T conditions. The mechanisms of their formation can be classified into several types, as described in Section 5.1, Section 5.2, Section 5.3, Section 5.4, Section 5.5 and Section 5.6.

5.1. Rapid Sedimentation

As depicted in Figure 7 and Figure 8a, continuous burial (with the seafloor shifting from Seafloor2 to Seafloor1) facilitates the transfer of deep heat to shallower, newly deposited strata. As the geothermal gradient is re-established, the GHSZ migrates upward from BGHSZ2 to BGHSZ1, leading to the formation of a new BSR in the shallower strata. Multiple rapid sedimentation events can induce repeated adjustments, resulting in the identification of double or multiple BSRs. Upward adjustments of the BGHSZ due to this mechanism have been observed in various marine regions, including the Black Sea, the Hikurangi subduction zone off New Zealand, the Nankai Trough in Japan, and the Manila subduction zone, as shown in Table 3, Table 5 and Table 7. Through numerical simulations, Gupta et al. (2024) identified the primary factors influencing the formation of multiple BSRs: continuous sedimentation layers several tens of meters thick, low-permeability sediment layers (which facilitate the accumulation of free gas), and appropriate kinetics of hydrate formation and dissociation—specifically, slower dissociation rates (10−16 mol/m2 Pa·s to 10−14 mol/m2 Pa·s) coupled with higher formation rates (10−14 mol/m2 Pa·s to 10−12 mol/m2 Pa·s) [222].

5.2. Tectonic Uplift and Subsidence

Figure 8b demonstrates that during tectonic uplift, a reduction in pressure drives the upward migration of the BGHSZ. As gas hydrates dissociate between the former BGHSZ and the newly established BGHSZ, two nearly parallel BSRs may temporarily coexist during this transition. If hydrates are present above the former BGHSZ, their dissociation during uplift can be significantly slowed due to the latent heat-buffering effect, thereby preserving BSR2 for a prolonged period [225]. It is important to note that the inferred uplift associated with double BSRs may result from multiple uplift events. A new BSR can only form if the quiescent period between successive uplift events is sufficiently long, allowing the dissociation of hydrates in the sediments above the former BGHSZ, the upward migration of free gas, and its subsequent accumulation beneath the newly established BGHSZ. Meanwhile, during tectonic subsidence, increased pressure drives the downward migration of the BGHSZ, in contrast to the process observed during tectonic uplift.

5.3. Upwelling of Thermal Fluids

According to Figure 8c, in regions of localized basement uplift, deep fluids migrate upward along faults, leading to an increased geothermal gradient in the shallow sediments. This results in the upward adjustment of the BGHSZ and the formation of anomalously shallow, arching BSRs [40]. This mechanism is similar to the BGHSZ migration induced by bottom water temperature (BWT) increases during interglacial periods [30,211,226]. However, unlike BWT-induced changes, the influence of thermal fluids on the original hydrate layer is typically more persistent and direct. As a result, this process does not necessarily lead to the formation of double or multiple BSRs.

5.4. Fluid Overpressure Induced by Slope Failure

As presented in Figure 8d, observations of microseismic activity and seismic attenuation indicate that highly overpressured fluids, generated at compressional margins, are expelled through the fracture network in the overlying strata following slope failure [227,228]. Shortly after slope failure, the upward migration of fluids increases pore pressure beneath the thrust ridge, causing the BGHSZ to shift downward [212]. In addition to slope failure, fluid overpressure can also result from various factors, including changes in fluid volume (e.g., temperature-induced expansion, fluid hydrocarbons generated from kerogen maturation, or H2O release during smectite-to-illite transformation), fluid movement (e.g., buoyancy-driven flow or permeability effects), and compaction (e.g., reservoir compaction induced by tectonic stress or rapid sedimentation) [229].

5.5. Canyon Erosion and Subsequent Sedimentation

As shown by Figure 8e, an increase in seafloor depth on the canyon’s eroded side leads to a downward shift of the BGHSZ. Conversely, on the channel deposition side, sediment accumulation tends to cause an upward migration of the BGHSZ. Between BGHSZ1 and BGHSZ2, residual hydrates and/or free gas may persist. Under these contrasting P-T conditions, a distinct intersecting double BSR can ultimately form [215,230].

5.6. Different Types of Gas Hydrate Structure

In the scenario illustrated by Figure 8f, the mechanism is relatively unique, as both BSR1 and BSR2 remain in equilibrium under current P-T conditions. Their formation is closely linked to the presence of different gas hydrate structures [30,208,209,224]. Heavy hydrocarbons preferentially form sII hydrates, which exhibit stability under higher P-T conditions compared to structure I (sI) hydrates [211]. Therefore, BSR2 indicates the lower boundary of the Type sII GHSZ, typically situated beneath BSR1, which marks the base of the Type SI GHSZ [219]. In the case study from the Norwegian continental margin, the occurrence of a double BSR is more reasonably explained by this theoretical model than by invoking the existence of a paleo-BSR [209].
In addition, another category of double/multiple BSRs is unrelated to gas hydrates but instead results from variations in sediment lithology or mineralogy [220,231]. However, since this category is beyond the scope of this study, it will not be further discussed here.
The interpretations of double/multiple BSRs described above are reasonable within specific regions but cannot be broadly generalized due to the complexity and variability of geological controlling factors worldwide. Consequently, research on double/multiple BSR phenomena remains exploratory. Studying double BSRs contributes to a better understanding of the evolutionary mechanisms, timescales, and distribution patterns of gas hydrate systems.

6. Challenges

As a key seismic indicator for identifying natural gas hydrates, the accurate interpretation of the BSR faces multiple technical challenges. Conventional P-wave seismic profiles often exhibit non-uniqueness in BSR morphology, making it difficult to reliably distinguish real BSRs from pseudo ones. While AVO analysis provides hydrocarbon indicators, its dependence on empirical relationships and indirect S-wave estimation introduces uncertainties in fluid characterization. Although multi-wave, multi-component techniques based on OBS enable in-situ S-wave measurements, their widespread application is constrained by high acquisition costs and complex data processing and interpretation. Therefore, economically acquiring reliable P- and S-wave data and establishing efficient interpretation workflows have become critical bottlenecks in seismic exploration of gas hydrates.
The fundamental challenge in BSR research lies in characterizing the coupled seismic responses between the overlying hydrate-bearing layer and the underlying free gas zone. While conventional P-wave seismic interpretation provides qualitative geometrical analysis of BSRs, it proves insufficient for comprehensive reservoir assessment. To address this limitation, the research community must develop integrated quantitative methodologies that incorporate ① joint inversion of elastic properties characterizing the gas hydrate-free gas system, ② mechanistic studies of seismic attenuation and dispersion, and ③ physics-based correlations between BSR attributes (both dynamic and static) and hydrate saturation through rigorous rock physics modeling. These advances will ultimately transform BSR analysis from simple phase-boundary identification to the quantitative inversion of reservoir physical parameters.

7. Conclusions

This study synthesizes global BSR observations to systematically investigate the origins of pseudo-reflections and double/multiple BSRs, the relationship between BSR and the base of the BGHSZ, and the primary controls on BSR seismic morphology. The following conclusions can be drawn:
  • BSRs, free gas zones, and GHBS exhibit complex, non-unique correlations. Reliable BSR identification requires integrating multiple seismic attributes rather than relying on a single diagnostic feature.
  • Over 35% of global BSRs occur significantly shallower than theoretical BGHSZ predictions, particularly in deepwater basins. This confirms that the BSR primarily marks the base of the gas hydrate occurrence zone (BGHOZ) and the top of the free gas zone (TFGZ), not necessarily the thermodynamic GHSZ boundary.
  • BSR morphology exhibits systematic variability controlled by three interlinked factors: ① geological setting, ② sediment properties, and ③ seismic acquisition constraints. Continuous BSRs typically develop in venting settings and in stratigraphic contexts where homogeneous strata are parallel to the BGHSZ. Discontinuous BSRs are often found in structural settings or where heterogeneous strata dip relative to the BGHSZ. Clustered BSR is predominantly associated with coarse-grained sediments in structural settings while pluming BSR is confined to venting settings linked to surface mud volcanoes or salt diapirs. The majority (~70–80%) of BSRs occur in fine-grained, grain-displacive sediments, where gas hydrates typically form as isolated lenses or nodules. In contrast, coarse-grained, pore-filling sediments account for a smaller proportion of BSR occurrences, typically less than 20%. Additionally, seismic data quality parameters—notably lateral resolution and survey-line orientation relative to geological structures—fundamentally govern the apparent continuity and sharpness of BSR reflections in seismic imagery.
  • Double and multiple BSRs result from dynamic adjustments in gas hydrate systems, driven by sedimentation, tectonics, fluid migration, or differing hydrate phase stability. While these mechanisms provide regional explanations, their complexity demands further interdisciplinary research to predict hydrate system evolution.
  • By integrating these four frontiers—frequency-dependent AVO, viscoelastic full waveform inversion (FWI), multicomponent anisotropy, and microstructural modeling—the next generation of BSR research will achieve quantitative, physics-constrained reservoir evaluation.
  • The integration of high-resolution exploration technologies and intelligent algorithms is anticipated to shift BSR research from “static feature identification” to “dynamic process analysis”, potentially optimizing gas hydrate exploration efficiency while providing critical scientific insights into the interaction mechanisms between hydrate systems and marine environments.

Author Contributions

Conceptualization, S.S.; methodology, S.S. and L.Z.; validation, L.Z. and H.L.; formal analysis, S.S.; investigation, S.S.; resources, W.C. and R.Y.; writing—original draft preparation, S.S.; writing—review and editing, L.Z. and H.L.; visualization, S.S.; supervision, L.Z. and H.L.; project administration, L.Z.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by DD20221703 and DD20230063 by the China Geological Survey and The Guangdong Major project of Basic and Applied Basic Research (No. 2020B0301030003).

Data Availability Statement

Data are contained within the article. The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AAmplitude
ACMActive Continental Margin
AVOAmplitude Versus Offset
BGHOZBase of the Hydrate Occurrence Zone
BGHSZBase of the Hydrate Stability Zone
BSRBottom-Simulating Reflector
CCContinental Collision
Clus.Clustered BSR
Cont.Continuous BSR
CSMHYDColorado School of Mines Hydrate
Diag.Diagenetic BSR
Disc.Discontinuous BSR
fDominant Seismic Frequency
FWIFull Waveform Inversion
GGGeothermal Gradient
GHBSGas Hydrate-bearing Sediment
GHSZGas Hydrate Stability Zone
GSTGeological Setting Types of BSRs
hVertical Resolution
ISBInland Sedimentary Basin
MCMMixed Continental Margin
NANot Available
PCMPassive Continental Margin
Plum.Pluming BSR
RReflection Coefficient
rHorizontal Resolution
sIStructure I
sIIStructure II
SMTSeismic Morphology Types of BSRs
Stra.Stratigraphic Setting
Stru.Structural Setting
TFGZTop of the Free Gas Zone
TGFTectonic Geological Features
TPBTransition Plate Boundary
TRLTectonic Rift Lake
TsfSeafloor Temperature
vSeismic Wave Velocity
Vent.Venting Setting
WDWater Depth
ZAcoustic Impedance
zTarget Depth
ZBGHSZDepth of BGHSZ in Salt Water (3.5 wt.% NaCl)
ZBGHSZ*Depth of BGHSZ in Pure Water
ZBSRBSR Depth
ρDensity
ΔZDepth Deviation between BGHSZ and BSR

References

  1. Trofimuk, A.A.; Cherskiy, N.V.; Makagon, Y.F.; Tsarev, V.P. Possible Mechanism of the Accumulation of Natural Gas. Int. Geol. Rev. 1973, 15, 1042–1046. [Google Scholar] [CrossRef]
  2. Stoll, R.D.; Ewing, J.; Bryan, G.M. Anomalous Wave Velocities in Sediments Containing Gas Hydrates. J. Geophys. Res. 1971, 76, 2090–2094. [Google Scholar] [CrossRef]
  3. Markl, R.G.; Bryan, G.M.; Ewing, J.I. Structure of the Blake-Bahama Outer Ridge. J. Geophys. Res. 1970, 75, 4539–4555. [Google Scholar] [CrossRef]
  4. Bangs, N.L.B.; Sawyer, D.S.; Golovchenko, X. Free Gas at the Base of the Gas Hydrate Zone in the Vicinity of the Chile Triple Junction. Geology 1993, 21, 905. [Google Scholar] [CrossRef]
  5. Langan, R.T.; Lerche, I.; Cutler, R.T. The Tracing of Rays Through Heterogeneous Media. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 8 April 1985. [Google Scholar]
  6. Singh, S.C.; Minshull, T.A.; Spence, G.D. Velocity Structure of a Gas Hydrate Reflector. Science 1993, 260, 204–207. [Google Scholar] [CrossRef]
  7. Singh, S.C.; Minshull, T.A. Velocity Structure of a Gas Hydrate Reflector at Ocean Drilling Program Site 889 from a Global Seismic Waveform Inversion. J. Geophys. Res. Solid Earth 1994, 99, 24221–24233. [Google Scholar] [CrossRef]
  8. Tarantola, A. A Strategy for Nonlinear Elastic Inversion of Seismic Reflection Data. Geophysics 1986, 51, 1893–1903. [Google Scholar] [CrossRef]
  9. Tarantola, A. Inversion of Seismic Reflection Data in the Acoustic Approximation. Geophysics 1984, 49, 1259–1266. [Google Scholar] [CrossRef]
  10. Hyndman, R.D.; Spence, G.D. A Seismic Study of Methane Hydrate Marine Bottom Simulating Reflectors. J. Geophys. Res. Solid Earth 1992, 97, 6683–6698. [Google Scholar] [CrossRef]
  11. Murphy, W.; Reischer, A.; Hsu, K. Modulus Decomposition of Compressional and Shear Velocities in Sand Bodies. Geophysics 1993, 58, 227–239. [Google Scholar] [CrossRef]
  12. Chand, S.; Minshull, T.A. Seismic Constraints on the Effects of Gas Hydrate on Sediment Physical Properties and Fluid Flow: A Review. Geofluids 2003, 3, 275–289. [Google Scholar] [CrossRef]
  13. Katzman, R.; Holbrook, W.S.; Paull, C.K. Combined Vertical-incidence and Wide-angle Seismic Study of a Gas Hydrate Zone, Blake Ridge. J. Geophys. Res. Solid Earth 1994, 99, 17975–17995. [Google Scholar] [CrossRef]
  14. Petersen, C.J.; Bünz, S.; Hustoft, S.; Mienert, J.; Klaeschen, D. High-Resolution P-Cable 3D Seismic Imaging of Gas Chimney Structures in Gas Hydrated Sediments of an Arctic Sediment Drift. Mar. Pet. Geol. 2010, 27, 1981–1994. [Google Scholar] [CrossRef]
  15. Lu, S.; McMechan, G.A. Elastic Impedance Inversion of Multichannel Seismic Data from Unconsolidated Sediments Containing Gas Hydrate and Free Gas. Geophysics 2004, 69, 164–179. [Google Scholar] [CrossRef]
  16. Lu, S.; McMechan, G.A. Estimation of Gas Hydrate and Free Gas Saturation, Concentration, and Distribution from Seismic Data. Geophysics 2002, 67, 582–593. [Google Scholar] [CrossRef]
  17. Bünz, S.; Mienert, J.; Vanneste, M.; Andreassen, K. Gas Hydrates at the Storegga Slide: Constraints from an Analysis of Multicomponent, Wide-Angle Seismic Data. Geophysics 2005, 70, B19–B34. [Google Scholar] [CrossRef]
  18. Ecker, C. Seismic Characterization of Methane Hydrate Structures. Ph.D. Thesis, Stanford University, San Francisco, CA, USA, 2001. [Google Scholar]
  19. Li, C.; Gou, L.; You, J.; Liu, X.; Ou, C. Further Studies on the Numerical Simulation of Bubble Plumes in the Cold Seepage Active Region. Acta Oceanol. Sin. 2016, 35, 118–124. [Google Scholar] [CrossRef]
  20. Westbrook, G.K.; Thatcher, K.E.; Rohling, E.J.; Piotrowski, A.M.; Pälike, H.; Osborne, A.H.; Nisbet, E.G.; Minshull, T.A.; Lanoisellé, M.; James, R.H.; et al. Escape of Methane Gas from the Seabed along the West Spitsbergen Continental Margin. Geophys. Res. Lett. 2009, 36, L15608. [Google Scholar] [CrossRef]
  21. Guerin, G.; Goldberg, D. Sonic Waveform Attenuation in Gas Hydrate-bearing Sediments from the Mallik 2L-38 Research Well, Mackenzie Delta, Canada. J. Geophys. Res. Solid Earth 2002, 107, EPM 1-1–EPM 1-11. [Google Scholar] [CrossRef]
  22. Matsushima, J. Seismic Attenuation from VSP Data in Methane Hydrate-Bearing Sediments. Explor. Geophys. 2007, 38, 29–36. [Google Scholar] [CrossRef]
  23. Li, C.; Feng, K.; Liu, X. Study on P-Wave Attenuation in Hydrate-Bearing Sediments Based on BISQ Model. J. Geol. Res. 2013, 2013, 176579. [Google Scholar] [CrossRef]
  24. Dewangan, P.; Mandal, R.; Jaiswal, P.; Ramprasad, T.; Sriram, G. Estimation of Seismic Attenuation of Gas Hydrate Bearing Sediments from Multi-Channel Seismic Data: A Case Study from Krishna–Godavari Offshore Basin. Mar. Pet. Geol. 2014, 58, 356–367. [Google Scholar] [CrossRef]
  25. Shipley, T.H.; Houston, H.M.; Buffler, R.T.; Shaub, F.J.; McMillen, K.J.; Ladd, J.W.; Worzel, J.L. Seismic Evidence for Widespread Possible Gas Hydrate Horizons on Continental Slopes and Rises. Am. Assoc. Pet. Geol. Bull. 1979, 63, 2204–2213. [Google Scholar] [CrossRef]
  26. Hyndman, R.D.; Davis, E.E. A Mechanism for the Formation of Methane Hydrate and Seafloor Bottom-simulating Reflectors by Vertical Fluid Expulsion. J. Geophys. Res. Solid Earth 1992, 97, 7025–7041. [Google Scholar] [CrossRef]
  27. Yun, T.S.; Francisca, F.M.; Santamarina, J.C.; Ruppel, C. Compressional and Shear Wave Velocities in Uncemented Sediment Containing Gas Hydrate. Geophys. Res. Lett. 2005, 32, L10609. [Google Scholar] [CrossRef]
  28. Rajan, A.; Mienert, J.; Bünz, S. Acoustic Evidence for a Gas Migration and Release System in Arctic Glaciated Continental Margins Offshore NW-Svalbard. Mar. Pet. Geol. 2012, 32, 36–49. [Google Scholar] [CrossRef]
  29. Bünz, S.; Polyanov, S.; Vadakkepuliyambatta, S.; Consolaro, C.; Mienert, J. Active Gas Venting through Hydrate-Bearing Sediments on the Vestnesa Ridge, Offshore W-Svalbard. Mar. Geol. 2012, 332–334, 189–197. [Google Scholar] [CrossRef]
  30. Bangs, N.L.B.; Musgrave, R.J.; Tréhu, A.M. Upward Shifts in the Southern Hydrate Ridge Gas Hydrate Stability Zone Following Postglacial Warming, Offshore Oregon. J. Geophys. Res. Solid Earth 2005, 110, B03102. [Google Scholar] [CrossRef]
  31. Berndt, C.; Bünz, S.; Clayton, T.; Mienert, J.; Saunders, M. Seismic Character of Bottom Simulating Reflectors: Examples from the Mid-Norwegian Margin. Mar. Pet. Geol. 2004, 21, 723–733. [Google Scholar] [CrossRef]
  32. Liang, J.-q.; Deng, W.; Lu, J.-a.; Kuang, Z.-g.; He, Y.-l.; Zhang, W.; Gong, Y.-h.; Liang, J.; Meng, M.-m. A Fast Identification Method Based on the Typical Geophysical Differences between Submarine Shallow Carbonates and Hydrate Bearing Sediments in the Northern South China Sea. China Geol. 2020, 3, 16–27. [Google Scholar] [CrossRef]
  33. Qian, J.; Kang, D.; Jin, J.; Lin, L.; Guo, Y.; Meng, M.; Wang, Z.; Wang, X. Quantitative Seismic Characterization for Gas Hydrate- and Free Gas-Bearing Sediments in the Shenhu Area, South China Sea. Mar. Pet. Geol. 2022, 139, 105606. [Google Scholar] [CrossRef]
  34. Spence, G.D.; Haacke, R.R.; Hyndman, R.D. 4. Seismic Indicators of Natural Gas Hydrate and Underlying Free Gas. In Geophysical Characterization of Gas Hydrates; Society of Exploration Geophysicists: Houston, TX, USA, 2010; pp. 39–71. [Google Scholar]
  35. Chopra, S.; Marfurt, K.J. Seismic Attributes—A Historical Perspective. Geophysics 2005, 70, 3SO–28SO. [Google Scholar] [CrossRef]
  36. Satyavani, N.; Sain, K.; Lall, M.; Kumar, B.J.P. Seismic Attribute Study for Gas Hydrates in the Andaman Offshore India. Mar. Geophys. Res. 2008, 29, 167–175. [Google Scholar] [CrossRef]
  37. Hornbach, M.J. Bottom Simulating Reflections Below the Blake Ridge, Western North Atlantic Margin. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 131–138. [Google Scholar]
  38. Hjelstuen, B.O.; Haflidason, H.; Sejrup, H.P.; Nygård, A. Sedimentary and Structural Control on Pockmark Development—Evidence from the Nyegga Pockmark Field, NW European Margin. Geo-Mar. Lett. 2010, 30, 221–230. [Google Scholar] [CrossRef]
  39. Vadakkepuliyambatta, S.; Chand, S.; Waage, M.; Bünz, S. Occurrence and Distribution of Bottom Simulating Reflections in the Barents Sea. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 237–245. [Google Scholar]
  40. Jin, J.; Wang, X.; Guo, Y.; Li, J.; Li, Y.; Zhang, X.; Qian, J.; Sun, L. Geological Controls on the Occurrence of Recently Formed Highly Concentrated Gas Hydrate Accumulations in the Shenhu Area, South China Sea. Mar. Pet. Geol. 2020, 116, 104294. [Google Scholar] [CrossRef]
  41. Portnov, A.; Cook, A.E.; Sawyer, D.E.; Yang, C.; Hillman, J.I.T.; Waite, W.F. Clustered BSRs: Evidence for Gas Hydrate-Bearing Turbidite Complexes in Folded Regions, Example from the Perdido Fold Belt, Northern Gulf of Mexico. Earth Planet. Sci. Lett. 2019, 528, 115843. [Google Scholar] [CrossRef]
  42. Liu, B.; Syed, W.H.; Chen, J.; Deng, X.; Yang, L.; Azevedo, L.; Duan, M.; Wu, T.; Ma, J.; Li, K. Distinct BSRs and Their Implications for Natural Gas Hydrate Formation and Distribution at the Submarine Makran Accretionary Zone. J. Oceanol. Limnol. 2021, 39, 1871–1886. [Google Scholar] [CrossRef]
  43. Shedd, W.; Boswell, R.; Frye, M.; Godfriaux, P.; Kramer, K. Occurrence and Nature of “Bottom Simulating Reflectors” in the Northern Gulf of Mexico. Mar. Pet. Geol. 2012, 34, 31–40. [Google Scholar] [CrossRef]
  44. Taner, M.T.; Koehler, F.; Sheriff, R.E. Complex Seismic Trace Analysis. Geophysics 1979, 44, 1041–1063. [Google Scholar] [CrossRef]
  45. Kim, K.J.; Yi, B.Y.; Kang, N.K.; Yoo, D.G. Seismic Attribute Analysis of the Indicator for Gas Hydrate Occurrence in the Northwest Ulleung Basin, East Sea. In Proceedings of the Energy Procedia; Elsevier Ltd.: Amsterdam, The Netherlands, 2015; Volume 76, pp. 463–469. [Google Scholar]
  46. Shankar, U.; Sain, K. Seismic Attributes for Characterization of Gas Hydrate and Free Gas in the Mahanadi Basin. In Proceedings of the 2015 Biennial International Conference & Exposition, Jaipur, India, 4–7 December 2015. [Google Scholar]
  47. Yang, R.; Yan, P.; Wu, N.; Sha, Z.; Liang, J. Application of AVO Analysis to Gas Hydrates Identification in the Northern Slope of the South China Sea. Acta Geophys. 2014, 62, 802–817. [Google Scholar] [CrossRef]
  48. Minshull, T.; White, R. Sediment Compaction and Fluid Migration in the Makran Accretionary Prism. J. Geophys. Res. Solid Earth 1989, 94, 7387–7402. [Google Scholar] [CrossRef]
  49. Lee, M.W.; Hutchinson, D.R.; Agena, W.F.; Dillon, W.P.; Miller, J.J.; Swift, B.A. Seismic Character of Gas Hydrates on the Southeastern U.S. Continental Margin. Mar. Geophys. Res. 1994, 16, 163–184. [Google Scholar] [CrossRef]
  50. Carcione, J.M.; Tinivella, U. Bottom-simulating Reflectors: Seismic Velocities and AVO Effects. Geophysics 2000, 65, 54–67. [Google Scholar] [CrossRef]
  51. Andreassen, K.; Hart, P.E.; MacKay, M. Amplitude versus Offset Modeling of the Bottom Simulating Reflection Associated with Submarine Gas Hydrates. Mar. Geol. 1997, 137, 25–40. [Google Scholar] [CrossRef]
  52. Treude, T.; Boetius, A.; Knittel, K.; Wallmann, K.; Barker Jørgensen, B. Anaerobic Oxidation of Methane above Gas Hydrates at Hydrate Ridge, NE Pacific Ocean. Mar. Ecol. Prog. Ser. 2003, 264, 1–14. [Google Scholar] [CrossRef]
  53. Depreiter, D.; Poort, J.; Van Rensbergen, P.; Henriet, J.P. Geophysical Evidence of Gas Hydrates in Shallow Submarine Mud Volcanoes on the Moroccan Margin. J. Geophys. Res. Solid Earth 2005, 110, B10103. [Google Scholar] [CrossRef]
  54. Charlou, J.L.; Donval, J.P.; Fouquet, Y.; Ondreas, H.; Knoery, J.; Cochonat, P.; Levaché, D.; Poirier, Y.; Jean-Baptiste, P.; Fourré, E.; et al. Physical and Chemical Characterization of Gas Hydrates and Associated Methane Plumes in the Congo–Angola Basin. Chem. Geol. 2004, 205, 405–425. [Google Scholar] [CrossRef]
  55. Hovland, M.T.; Roy, S. Shallow Gas Hydrates Near 64° N, Off Mid-Norway: Concerns Regarding Drilling and Production Technologies. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 15–32. [Google Scholar]
  56. Ketzer, M.; Viana, A.; Miller, D.; Augustin, A.; Rodrigues, F.; Praeg, D.; Cupertino, J.; Freire, F.; Kowsmann, R.; Dickens, G.R.; et al. Gas Hydrate Systems on the Brazilian Continental Margin. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 343–352. [Google Scholar]
  57. Bangs, N.L.; Johnson, J.E.; Tréhu, A.M.; Arsenault, M.A. Hydrate Ridge—A Gas Hydrate System in a Subduction Zone Setting. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 89–107. [Google Scholar]
  58. Berndt, C.; Davies, R.; Li, A.; Yang, J. Insights into Gas Hydrate Dynamics from 3D Seismic Data, Offshore Mauritania. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 323–329. [Google Scholar]
  59. Kvenvolden, K.A.; Kaster, M. Gas Hydrates of the Peruvian Outer Continental Margin. In Proceedings of the Ocean Drilling Program, 112 Scientific Reports; Ocean Drilling Program: College Station, TX, USA, 1990. [Google Scholar]
  60. Maselli, V.; Iacopini, D.; Ahaneku, C.V.; Micallef, A.; Green, A. First Evidence of Bottom Simulation Reflectors in the Western Indian Ocean Offshore Tanzania. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 427–433. [Google Scholar]
  61. Kenyon, N.H.; Ivanov, M.K.; Akhmetzhanov, A.M.; Akhmanov, G.G. Multidisciplinary Study of Geological Processes on the North East Atlantic and Western Mediterranean Margins: Preliminary Results of Geological and Geophysical Investigations during the TTR-9 Cruise of R/V Professor Logachev June-July, 1999; IOC. Technical Series; UNESCO: London, UK, 2000; Volume 56. [Google Scholar]
  62. Somoza, L.; Medialdea, T.; Gonzalez, F.J. Bottom Simulating Reflectors Along the Scan Basin, a Deep-Sea Gateway Between the Weddell Sea (Antarctica) and Scotia Sea. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 483–492. [Google Scholar]
  63. Boyce, R.E. Leg XI Grain Size Analysis. In Initial Reports of the Deep Sea Drilling Project, 11; U.S. Government Printing Office: Washington, DC, USA, 1972. [Google Scholar]
  64. Sheridan, R.E. Site 533: Blake Outer Ridge. In Initial Reports of the Deep Sea Drilling Project, 76; U.S. Government Printing Office: Washington, DC, USA, 1983; Volume 76, pp. 35–140. [Google Scholar]
  65. Collett, T.S.; Wendlandt, R.F. Formation Evaluation of Gas Hydrate-Bearing Marine Sediments on the Blake Ridge with Downhole Geochemical Log Measurements. In Proceedings of the Ocean Drilling Program, 164 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 2000. [Google Scholar]
  66. Matsumoto, R.; Uchida, T.; Waseda, A.; Uchida, T.; Takeya, S.; Hirano, T.; Yamada, K.; Maeda, Y.; Okui, T. Occurrence, Structure, and Composition of Natural Gas Hydrate Recovered from the Blake Ridge, Northwest Atlantic. In Proceedings of the Ocean Drilling Program, 164 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 2000. [Google Scholar]
  67. Paull, C.K.; Matsumoto, R.; Wallace, P.J. Site 997. In Proceedings of the Ocean Drilling Program, 164 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1996; Volume 164, pp. 277–334. [Google Scholar]
  68. Paull, C.K.; Matsumoto, R.; Wallace, P.J. Site 995. In Proceedings of the Ocean Drilling Program, 164 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1996; Volume 164, pp. 175–240. [Google Scholar]
  69. Hillman, J.I.T.; Cook, A.E.; Sawyer, D.E.; Küçük, H.M.; Goldberg, D.S. The Character and Amplitude of ‘Discontinuous’ Bottom-Simulating Reflections in Marine Seismic Data. Earth Planet. Sci. Lett. 2017, 459, 157–169. [Google Scholar] [CrossRef]
  70. Vanneste, M.; De Batist, M.; Golmshtok, A.; Kremlev, A.; Versteeg, W. Multi-Frequency Seismic Study of Gas Hydrate-Bearing Sediments in Lake Baikal, Siberia. Mar. Geol. 2001, 172, 1–21. [Google Scholar] [CrossRef]
  71. Fohrmann, M.; Pecher, I.A. Analysing Sand-Dominated Channel Systems for Potential Gas-Hydrate-Reservoirs Using an AVO Seismic Inversion Technique on the Southern Hikurangi Margin, New Zealand. Mar. Pet. Geol. 2012, 38, 19–34. [Google Scholar] [CrossRef]
  72. Waghorn, K.A.; Bünz, S.; Plaza-Faverola, A.; Johnson, J.E. 3-D Seismic Investigation of a Gas Hydrate and Fluid Flow System on an Active Mid-Ocean Ridge; Svyatogor Ridge, Fram Strait. Geochem. Geophys. Geosyst. 2018, 19, 2325–2341. [Google Scholar] [CrossRef]
  73. Johnson, J.E.; Mienert, J.; Plaza-Faverola, A.; Vadakkepuliyambatta, S.; Knies, J.; Bünz, S.; Andreassen, K.; Ferré, B. Abiotic Methane from Ultraslow-Spreading Ridges Can Charge Arctic Gas Hydrates. Geology 2015, 43, 371–374. [Google Scholar] [CrossRef]
  74. Lin, C.-C.; Lin, A.T.-S.; Liu, C.-S.; Horng, C.-S.; Chen, G.-Y.; Wang, Y. Canyon-Infilling and Gas Hydrate Occurrences in the Frontal Fold of the Offshore Accretionary Wedge off Southern Taiwan. Mar. Geophys. Res. 2014, 35, 21–35. [Google Scholar] [CrossRef]
  75. Berndt, C.; Chi, W.-C.; Jegen, M.; Lebas, E.; Crutchley, G.; Muff, S.; Hölz, S.; Sommer, M.; Lin, S.; Liu, C.-S.; et al. Tectonic Controls on Gas Hydrate Distribution Off SW Taiwan. J. Geophys. Res. Solid Earth 2019, 124, 1164–1184. [Google Scholar] [CrossRef]
  76. Leslie, S.C.; Mann, P. Distribution and Character of Bottom Simulating Reflections in the Western Caribbean Offshore Guajira Peninsula, Colombia. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 333–341. [Google Scholar]
  77. Portnov, A.; Cook, A.E.; Sawyer, D.E. Bottom Simulating Reflections and Seismic Phase Reversals in the Gulf of Mexico. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 315–322. [Google Scholar]
  78. Manley, P.L.; Pirmez, C.; Busch, W.; Cramp, A. Grain-Size Characterization of Amazon Fan Deposits and Comparison to Seismic Facies Units. In Proceedings of the Ocean Drilling Program, 155 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 1997. [Google Scholar]
  79. Manley, P.L.; Flood, R.D. Cyclic Sediment Deposition within Amazon Deep-Sea Fan. Am. Assoc. Pet. Geol. Bull. 1988, 72, 912–925. [Google Scholar] [CrossRef]
  80. Yun, T.S.; Narsilio, G.A.; Carlos Santamarina, J. Physical Characterization of Core Samples Recovered from Gulf of Mexico. Mar. Pet. Geol. 2006, 23, 893–900. [Google Scholar] [CrossRef]
  81. Pflaum, R.C.; Brooks, J.M.; Cox, H.B.; Kennicutt, M.C.; II, S. Molecular and Isotopic Analysis of Core Gases and Gas Hydrates, Deep Sea Drilling Project Leg 96. In Initial Reports of the Deep Sea Drilling Project, 96; U.S. Government Printing Office: Washington, DC, USA, 1986. [Google Scholar]
  82. Paull, C.K.; Shipboard Scientific Party. Sites 991/992/993. In Proceedings of the Ocean Drilling Program, 164 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1996; Volume 164, pp. 65–97. [Google Scholar]
  83. Uchida, T.; Yamamoto, J.; Okada, S.; Waseda, A.; Baba, K.; Okatsu, K.; Matsumoto, R. Shipbord Scientific Party of ODP Leg 164. In Methane Hydrates in Deep Marine Sediments—X-Ray CT and NMR Studies of ODP Leg 164 Hydrates; Jitsugyo Kohosha, 1997; pp. 36–42. [Google Scholar]
  84. Dillon, W.P.; Popenoe, P.; Grow, J.A.; Klitgord, K.D.; Swift, B.A.; Paull, C.K.; Cashman, K.V. Growth Faulting and Salt Diapirism: Their Relationship and Control in the Carolina Trough, Eastern North America/Subtitle. In Studies in Continental Margin Geology; American Association of Petroleum Geologists: Tulsa, OK, USA, 1982. [Google Scholar]
  85. Hollister, C.D.; Ewing, J.I.; Habib, D.; Hathaway, J.C.; Lancelot, Y.; Luterbacher, H.; Paulus, F.J.; Poag, C.W.; Wilcoxon, J.A.; Worstell, P. Site 107: Upper Continental Rise. In Initial Reports of the Deep Sea Drilling Project, 11; U.S. Government Printing Office: Washington, DC, USA, 1972. [Google Scholar]
  86. Brian, E.T.; Bryan, G.B.; Ewing, J.I. Gas-Hydrate Horizons Detected in Seismic-Profiler Data from the Western North Atlantic. Am. Assoc. Pet. Geol. Bull. 1977, 61, 698–707. [Google Scholar] [CrossRef]
  87. Wefer, G.; Berger, W.H.; Richter, C. Site 1080. In Proceedings of the Ocean Drilling Program 175 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1998; Volume 175, pp. 201–221. [Google Scholar]
  88. Swart, R. Hydrate Occurrences in the Namibe Basin, Offshore Namibia. In Sediment-Hosted Gas Hydrates: New Insights on Natural and Synthetic Systems; The Geological Society of London: London, UK, 2009; Volume 319, pp. 73–80. [Google Scholar] [CrossRef]
  89. de Santa Ana, H.; Latrónica, L.; Tomasini, J.; Morales, E.; Ferro, S.; Gristo, P.; Machado, L.; Veroslavsky, G.; Ucha, N. Economic and Exploratory Review of Gas Hydrates and Other Gas Manifestations of the Uruguayan Continental Shelf; University of British Columbia Library: Vancouver, BC, Canada, 2008. [Google Scholar]
  90. Pittenger, A.; Taylor, E.; Bryant, W.R. The Influence of Biogenic Silica on the Geotechnical Stratigraphy of the Vøring Plateau, Norwegian Sea. In Proceedings of the Ocean Drilling Program, 104 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 1989. [Google Scholar]
  91. Ginsburg, G.D.; Milkov, A.V.; Soloviev, V.A.; Egorov, A.V.; Cherkashev, G.A.; Vogt, P.R.; Crane, K.; Lorenson, T.D.; Khutorskoy, M.D. Gas Hydrate Accumulation at the Håkon Mosby Mud Volcano. Geo-Mar. Lett. 1999, 19, 57–67. [Google Scholar] [CrossRef]
  92. Senger, K.; Bünz, S.; Mienert, J. First-Order Estimation of In-Place Gas Resources at the Nyegga Gas Hydrate Prospect, Norwegian Sea. Energies 2010, 3, 2001–2026. [Google Scholar] [CrossRef]
  93. Mosher, D.C. A Margin-Wide BSR Gas Hydrate Assessment: Canada’s Atlantic Margin. Mar. Pet. Geol. 2011, 28, 1540–1553. [Google Scholar] [CrossRef]
  94. Mosher, D.C.; Hawken, J.E.; Campbell, D.C. Gas Hydrates and Submarine Sediment Mass Failure: A Case Study from Sackville Spur, Offshore Newfoundland. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 303–314. [Google Scholar]
  95. Hawken, J. Origin and Stratigraphic Setting of the Sackville Spur Bottom Simulating Reflector, Offshore Newfoundland. 2017. Available online: https://dalspace.library.dal.ca/items/ef6d5d69-fd2f-4a1a-8fa5-cdb6c8246b14 (accessed on 1 January 2020).
  96. Davies, R.J.; Yang, J.; Li, A.; Mathias, S.; Hobbs, R. An Irregular Feather-Edge and Potential Outcrop of Marine Gas Hydrate along the Mauritanian Margin. Earth Planet. Sci. Lett. 2015, 423, 202–209. [Google Scholar] [CrossRef]
  97. Warren, L.P. Site 502: Colombia Basin, Western Caribbean Sea. In Initial Reports of the Deep Sea Drilling Project, 68; U.S. Government Printing Office: Washington, DC, USA, 1982; Volume 68, pp. 15–162. [Google Scholar]
  98. Mayer, L.A. Physical Properties of Sediment Recovered on Deep Sea Drilling Project Leg 68 with the Hydraulic Piston Corer. In Initial Reports of the Deep Sea Drilling Project, 68; U.S. Government Printing Office: Washington, DC, USA, 1982; Volume 68, pp. 365–382. [Google Scholar]
  99. Reed, D.L.; Silver, E.A.; Tagudin, J.E.; Shipley, T.H.; Vrolijk, P. Relations between Mud Volcanoes, Thrust Deformation, Slope Sedimentation, and Gas Hydrate, Offshore North Panama. Mar. Pet. Geol. 1990, 7, 44–54. [Google Scholar] [CrossRef]
  100. Boyce, R.E. Grain Size Analysis, Leg 9. In Initial Reports of the Deep Sea Drilling Project, 9; U.S. Government Printing Office: Washington, DC, USA, 1972; Volume 9, pp. 779–796. [Google Scholar]
  101. Shipley, T.H.; Houston, H.H.; Buffler, R.T. Widespread Occurrence of Possible Gas-Hydrate Horizons from Continental Slopes as Identified on Seismic Reflection Profiles. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 8 April 1979. [Google Scholar]
  102. Mazurenko, L.L.; Soloviev, V.A.; Belenkaya, I.; Ivanov, M.K.; Pinheiro, L.M. Mud Volcano Gas Hydrates in the Gulf of Cadiz. Terra Nova 2002, 14, 321–329. [Google Scholar] [CrossRef]
  103. Wefer, G. Site 1076. In Proceedings of the Ocean Drilling Program 175 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1998; Volume 175, pp. 87–113. [Google Scholar]
  104. Pape, T.; Bohrmann, G. Shallow Gas Hydrates Associated to Pockmarks in the Northern Congo Deep-Sea Fan, SW Africa. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 359–371. [Google Scholar]
  105. Wenau, S.; Spieß, V.; Pape, T.; Fekete, N. Controlling Mechanisms of Giant Deep Water Pockmarks in the Lower Congo Basin. Mar. Pet. Geol. 2017, 83, 140–157. [Google Scholar] [CrossRef]
  106. Sultan, N.; Garziglia, S.; Ruffine, L. New Insights into the Transport Processes Controlling the Sulfate-Methane-Transition-Zone near Methane Vents. Sci. Rep. 2016, 6, 26701. [Google Scholar] [CrossRef]
  107. Wei, J.; Pape, T.; Sultan, N.; Colliat, J.-L.; Himmler, T.; Ruffine, L.; de Prunelé, A.; Dennielou, B.; Garziglia, S.; Marsset, T.; et al. Gas Hydrate Distributions in Sediments of Pockmarks from the Nigerian Margin—Results and Interpretation from Shallow Drilling. Mar. Pet. Geol. 2015, 59, 359–370. [Google Scholar] [CrossRef]
  108. Hovland, M.; Gallagher, J.W.; Clennell, M.B.; Lekvam, K. Gas Hydrate and Free Gas Volumes in Marine Sediments: Example from the Niger Delta Front. Mar. Pet. Geol. 1997, 14, 245–255. [Google Scholar] [CrossRef]
  109. Naim, F.; Cook, A.E. Gas Hydrate in the Faroe-Shetland Basin, Offshore United Kingdom. Mar. Pet. Geol. 2025, 176, 107347. [Google Scholar] [CrossRef]
  110. Moore, J.C.; Watkins, J.S.; Bachman, S.B.; Beghtel, F.W.; Didyk, B.M.; Leggett, J.K. Site 491. In Initial Reports of the Deep Sea Drilling Project, 66; U.S. Government Printing Office: Washington, DC, USA, 1982; Volume 66, pp. 219–287. [Google Scholar] [CrossRef]
  111. Tan, B.; Germaine, J.T.; Flemings, P.B. Data Report: Consolidation and Strength Characteristics of Sediments from ODP Site 1244, Hydrate Ridge, Cascadia Continental Margin. In Proceedings of the Ocean Drilling Program, 199 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 2006. [Google Scholar]
  112. Tréhu, A.M.; Bohrmann, G.; Rack, F.R.; Torres, M.E.; Shipboard Scientific Party. Site 1245. In Proceedings of the Ocean Drilling Program, 204 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 2003; Volume 204, pp. 1–31. [Google Scholar]
  113. Camerlenghi, A.; Lucchi, R.G.; Rothwell, R.G. Grain-Size Analysis and Distribution in Cascadia Margin Sediments, Northeastern Pacific. In Proceedings of the Ocean Drilling Program, 146 Part 1 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 1995. [Google Scholar]
  114. Westbrook, G.K. Sites 889 and 890. In Proceedings of the Ocean Drilling Program, 146 Part 1 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1994; Volume 146, pp. 127–239. [Google Scholar]
  115. Mazumdar, A.; João, H.M.; Peketi, A.; Dewangan, P.; Kocherla, M.; Joshi, R.K.; Ramprasad, T. Geochemical and Geological Constraints on the Composition of Marine Sediment Pore Fluid: Possible Link to Gas Hydrate Deposits. Mar. Pet. Geol. 2012, 38, 35–52. [Google Scholar] [CrossRef]
  116. Suess, E. Site 688. In Proceedings of the Ocean Drilling Program, 112 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1988; Volume 112, pp. 873–1004. [Google Scholar]
  117. Dai, S.; Santamarina, J.C.; Waite, W.F.; Kneafsey, T.J. Hydrate Morphology: Physical Properties of Sands with Patchy Hydrate Saturation. J. Geophys. Res. Solid Earth 2012, 117, B11205. [Google Scholar] [CrossRef]
  118. Tsuboi, S.; Kagami, H.; Tokuyama, H. Stratigraphy and Tectonic Interpretations of Multichannel Seismic Reflection Records across the Japan Trench off Sanriku, Northeastern Japan. In Initial Reports of the Deep Sea Drilling Project, 87; U.S. Government Printing Office: Washington, DC, USA, 1986; pp. 745–750. [Google Scholar]
  119. Uchida, T.; Lu, H.; Tomaru, H.; Matsumoto, R.; Senoh, O.; Oda, H.; Okada, S.; Delwiche, M.; Dallimore, S.R. Subsurface Occurrence of Natural Gas Hydrate in the Nankai Trough Area: Implication for Gas Hydrate Concentration. Resour. Geol. 2004, 54, 35–44. [Google Scholar] [CrossRef]
  120. Kinoshita, H.; Yamano, M. The Heat Flow Anomaly in the Nankai Trough Area. In Initial Reports of the Deep Sea Drilling Project, 87; U.S. Government Printing Office: Washington, DC, USA, 1986. [Google Scholar]
  121. Dang, H.; Luan, X.-W.; Chen, R.; Zhang, X.; Guo, L.; Klotz, M.G. Diversity, Abundance and Distribution of AmoA-Encoding Archaea in Deep-Sea Methane Seep Sediments of the Okhotsk Sea. FEMS Microbiol. Ecol. 2010, 72, 370–385. [Google Scholar] [CrossRef] [PubMed]
  122. Shoji, H.; Minami, H.; Hachikubo, A.; Sakagami, H.; Hyakutake, K.; Soloviev, V.; Matveeva, T.; Mazurenko, L.; Kaulio, V.; Gladysch, V.; et al. Hydrate-bearing Structures in the Sea of Okhotsk. Eos Trans. Am. Geophys. Union 2005, 86, 13–18. [Google Scholar] [CrossRef]
  123. Luan, X.; Jin, Y.; Obzhirov, A.; Yue, B. Characteristics of Shallow Gas Hydrate in Okhotsk Sea. Sci. China Ser. D Earth Sci. 2008, 51, 415–421. [Google Scholar] [CrossRef]
  124. Lüdmann, T.; Wong, H.K. Characteristics of Gas Hydrate Occurrences Associated with Mud Diapirism and Gas Escape Structures in the Northwestern Sea of Okhotsk. Mar. Geol. 2003, 201, 269–286. [Google Scholar] [CrossRef]
  125. Matveeva, T.; Logvina, E.; Nazarova, O.; Gladysh, V. Gas Hydrate-Bearing Province off Eastern Sakhalin Slope (Sea of Okhotsk): Geological Setting and Factors of Control. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 375–389. [Google Scholar]
  126. Kim, Y.-G.; Lee, S.-M.; Jin, Y.K.; Baranov, B.; Obzhirov, A.; Salomatin, A.; Shoji, H. The Stability of Gas Hydrate Field in the Northeastern Continental Slope of Sakhalin Island, Sea of Okhotsk, as Inferred from Analysis of Heat Flow Data and Its Implications for Slope Failures. Mar. Pet. Geol. 2013, 45, 198–207. [Google Scholar] [CrossRef]
  127. Lee, C.; Yun, T.S.; Lee, J.-S.; Bahk, J.J.; Santamarina, J.C. Geotechnical Characterization of Marine Sediments in the Ulleung Basin, East Sea. Eng. Geol. 2011, 117, 151–158. [Google Scholar] [CrossRef]
  128. Kim, G.Y.; Yi, B.Y.; Yoo, D.G.; Ryu, B.J.; Riedel, M. Evidence of Gas Hydrate from Downhole Logging Data in the Ulleung Basin, East Sea. Mar. Pet. Geol. 2011, 28, 1979–1985. [Google Scholar] [CrossRef]
  129. Wang, X.; Hutchinson, D.R.; Wu, S.; Yang, S.; Guo, Y. Elevated Gas Hydrate Saturation within Silt and Silty Clay Sediments in the Shenhu Area, South China Sea. J. Geophys. Res. 2011, 116, B05102. [Google Scholar] [CrossRef]
  130. Wang, X.; Collett, T.S.; Lee, M.W.; Yang, S.; Guo, Y.; Wu, S. Geological Controls on the Occurrence of Gas Hydrate from Core, Downhole Log, and Seismic Data in the Shenhu Area, South China Sea. Mar. Geol. 2014, 357, 272–292. [Google Scholar] [CrossRef]
  131. Liu, C.; Meng, Q.; He, X.; Li, C.; Ye, Y.; Zhang, G.; Liang, J. Characterization of Natural Gas Hydrate Recovered from Pearl River Mouth Basin in South China Sea. Mar. Pet. Geol. 2015, 61, 14–21. [Google Scholar] [CrossRef]
  132. Carter, R.; Mccave, I.N. Ocean Drilling Program Leg 181 Preliminary Report Southwest Pacific Gateways. In Proceedings of the Ocean Drilling Program; Ocean Drilling Program: College Station, TX, USA, 1998; Volume 181, ISBN 1223333450. [Google Scholar]
  133. Schwalenberg, K.; Haeckel, M.; Poort, J.; Jegen, M. Evaluation of Gas Hydrate Deposits in an Active Seep Area Using Marine Controlled Source Electromagnetics: Results from Opouawe Bank, Hikurangi Margin, New Zealand. Mar. Geol. 2010, 272, 79–88. [Google Scholar] [CrossRef]
  134. Pecher, I.; Crutchley, G.; Kröger, K.F.; Hillman, J.; Mountjoy, J.; Coffin, R.; Gorman, A. New Zealand’s Gas Hydrate Systems. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 415–424. [Google Scholar]
  135. Barnes, P.M.; Pecher, I.A.; LeVay, L.J.; Bourlange, S.M.; Brunet, M.M.Y.; Cardona, S.; Clennell, M.B.; Cook, A.E.; Crundwell, M.P.; Dugan, B.; et al. Site U1517; Texas A&M University: Houston, TX, USA, 2019. [Google Scholar]
  136. Brown, K.M.; Bangs, N.L.; Froelich, P.N.; Kvenvolden, K.A. The Nature, Distribution, and Origin of Gas Hydrate in the Chile Triple Junction Region. Earth Planet. Sci. Lett. 1996, 139, 471–483. [Google Scholar] [CrossRef]
  137. Behrmann, J.H.; Lewis, S.D.; Musgrave, R.J. Site 860. In Proceedings of the Ocean Drilling Program, 141 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1992; Volume 141, pp. 159–238. [Google Scholar]
  138. Bangs, N.L.B.; Sawyer, D.S.; Golovchenko, X. The Cause of the Bottom-Simulating Reflection in the Vicinity of the Chile Triple Junction. In Proceedings of the Ocean Drilling Program, 141 Scientific Results; Ocean Drilling Program: College Station, TX, USA, 1995. [Google Scholar]
  139. Collett, T.S.; Dallimore, S.R. Permafrost-Associated Gas Hydrate; Springer: Dordrecht, The Netherlands, 2003; pp. 43–60. [Google Scholar] [CrossRef]
  140. Chung, J.S.; Nativig, B.J.; Das, B.M. The Current Distribution and Thermal Stability of Natural Gas Hydrates in the Canadian Polar Regions; International Society of Offshore and Polar Engineers: Mountain View, CA, USA, 1994; ISBN 1880653109. [Google Scholar]
  141. Majorowicz, J.A.; Hannigan, P.K.; Osadetz, K.G. Study of the Natural Gas Hydrate “Trap Zone” and TheMethane Hydrate Potential in the Sverdrup Basin, Canada. Nat. Resour. Res. 2002, 11, 79–96. [Google Scholar] [CrossRef]
  142. Jansen, E. Site 986. In Proceedings of the Ocean Drilling Program, 162 Initial Reports; Ocean Drilling Program: College Station, TX, USA, 1996; Volume 162, pp. 287–343. [Google Scholar]
  143. Hustoft, S.; Bünz, S.; Mienert, J.; Chand, S. Gas Hydrate Reservoir and Active Methane-Venting Province in Sediments on <20 Ma Young Oceanic Crust in the Fram Strait, Offshore NW-Svalbard. Earth Planet. Sci. Lett. 2009, 284, 12–24. [Google Scholar] [CrossRef]
  144. Scholl, D.W.; Creager, J.S.; Boyce, R.E.; Echols, R.J.; The Shipboard Scientific Party. Site 185. In Initial Reports of the Deep Sea Drilling Project, 19; U.S. Government Printing Office: Washington, DC, USA, 1973; Volume 19, pp. 169–216. [Google Scholar]
  145. Ruppel, C.D.; Hart, P.E. Gas Hydrates on Alaskan Marine Margins. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 209–223. [Google Scholar]
  146. Bode, G.W. Grain Size. In Initial Reports of the Deep Sea Drilling Project, 19; U.S. Government Printing Office: Washington, DC, USA, 1973. [Google Scholar]
  147. Cooper, A.K.; Scholl, D.W.; Marlow, M.S.; Childs, J.R.; Redden, G.D.; Kvenvolden, K.A.; Stevenson, A.J. Hydrocarbon Potential of Aleutian Basin, Bering Sea. Am. Assoc. Pet. Geol. Bull. 1979, 63, 2070–2087. [Google Scholar] [CrossRef]
  148. Sættem, J.; Poole, D.A.R.; Ellingsen, L.; Sejrup, H.P. Glacial Geology of Outer Bjørnøyrenna, Southwestern Barents Sea. Mar. Geol. 1992, 103, 15–51. [Google Scholar] [CrossRef]
  149. Waage, M.; Portnov, A.; Serov, P.; Bünz, S.; Waghorn, K.A.; Vadakkepuliyambatta, S.; Mienert, J.; Andreassen, K. Geological Controls on Fluid Flow and Gas Hydrate Pingo Development on the Barents Sea Margin. Geochem. Geophys. Geosyst. 2019, 20, 630–650. [Google Scholar] [CrossRef]
  150. Andreassen, K.; Hogstad, K.; Berteussen, K.A. Gas Hydrate in the Southern Barents Sea, Indicated by a Shallow Seismic Anomaly. First Break 1990, 8, 235–245. [Google Scholar] [CrossRef]
  151. Cox, D.R.; Huuse, M.; Newton, A.M.W.; Sarkar, A.D.; Knutz, P.C. Shallow Gas and Gas Hydrate Occurrences on the Northwest Greenland Shelf Margin. Mar. Geol. 2021, 432, 106382. [Google Scholar] [CrossRef]
  152. Hayes, D.E. Site 273. In Initial Reports of the Deep Sea Drilling Project, 28; U.S. Government Printing Office: Washington, DC, USA, 1975; Volume 28, pp. 335–367. [Google Scholar]
  153. Geletti, R.; Busetti, M. A Double Bottom Simulating Reflector in the Western Ross Sea, Antarctica. J. Geophys. Res. Solid Earth 2011, 116, B04101. [Google Scholar] [CrossRef]
  154. Lonsdale, M.J. The Relationship between Silica Diagenesis|Methane|Seismic Reflections on the South Orkney Microcontinent. In Proceedings of the Ocean Drilling Program, 113 Scientific Reports; Ocean Drilling Program: College Station, TX, USA, 1990. [Google Scholar]
  155. The Shipboard Scientific Party. Site 222. In Initial Reports of the Deep Sea Drilling Project, 23; U.S. Government Printing Office: Washington, DC, USA, 1974. [Google Scholar]
  156. White, R.S. Gas Hydrate Layers Trapping Free Gas in the Gulf of Oman. Earth Planet. Sci. Lett. 1979, 42, 114–120. [Google Scholar] [CrossRef]
  157. Yun, T.S.; Fratta, D.; Santamarina, J.C. Hydrate-Bearing Sediments from the Krishna−Godavari Basin: Physical Characterization, Pressure Core Testing, and Scaled Production Monitoring. Energy Fuels 2010, 24, 5972–5983. [Google Scholar] [CrossRef]
  158. Collett, T.S.; Riedel, M.; Boswell, R.; Presley, J.; Kumar, P.; Sathe, A.; Sethi, A.; Lall, M.V. Indian National Gas Hydrate Program Expedition 01 Report; U.S. Geological Survey: Reston, VA, USA, 2015.
  159. Winters, W.J.; Waite, W.F.; Mason, D.H.; Kumar, P. Physical Properties of Repressurized Samples Recovered during the 2006 National Gas Hydrate Program Expedition Offshore India; University of British Columbia: Vancouver, BC, Canada, 2008. [Google Scholar]
  160. Kopp, H. BSR Occurrence along the Sunda Margin: Evidence from Seismic Data. Earth Planet. Sci. Lett. 2002, 197, 225–235. [Google Scholar] [CrossRef]
  161. Naim, F.; Cook, A.E. Gas Hydrate in the North Carnarvon Basin, Offshore Western Australia. Mar. Pet. Geol. 2024, 164, 106807. [Google Scholar] [CrossRef]
  162. Trimonis, E.S.; Shimkus, K.M. Grain-Size of the Black Sea Sediments, DSDP Leg 42B. In Initial Reports of the Deep Sea Drilling Project, 42 Pt. 2; U.S. Government Printing Office: Washington, DC, USA, 1978. [Google Scholar]
  163. Khlystov, O.; De Batist, M.; Shoji, H.; Hachikubo, A.; Nishio, S.; Naudts, L.; Poort, J.; Khabuev, A.; Belousov, O.; Manakov, A.; et al. Gas Hydrate of Lake Baikal: Discovery and Varieties. J. Asian Earth Sci. 2013, 62, 162–166. [Google Scholar] [CrossRef]
  164. Bialas, J.; Haeckel, M. Gas Hydrate Accumulations in the Black Sea. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 451–461. [Google Scholar]
  165. Zander, T.; Haeckel, M.; Berndt, C.; Chi, W.C.; Klaucke, I.; Bialas, J.; Klaeschen, D.; Koch, S.; Atgın, O. On the Origin of Multiple BSRs in the Danube Deep-Sea Fan, Black Sea. Earth Planet. Sci. Lett. 2017, 462, 15–25. [Google Scholar] [CrossRef]
  166. Kataoka, S.; Yamashita, S.; Kawaguchi, T.; Suzuki, T. The Soil Properties of Lake-Bottom Sediments in the Lake Baikal Gas Hydrate Province. Soils Found. 2009, 49, 757–775. [Google Scholar] [CrossRef]
  167. Lykousis, V.; Alexandri, S.; Woodside, J.; de Lange, G.; Dählmann, A.; Perissoratis, C.; Heeschen, K.; Ioakim, C.; Sakellariou, D.; Nomikou, P.; et al. Mud Volcanoes and Gas Hydrates in the Anaximander Mountains (Eastern Mediterranean Sea). Mar. Pet. Geol. 2009, 26, 854–872. [Google Scholar] [CrossRef]
  168. Praeg, D.; Migeon, S.; Mascle, J.; Unnithan, V.; Ketzer, M. A Gas Hydrate System of Heterogeneous Character in the Nile Deep-Sea Fan. In World Atlas of Submarine Gas Hydrates in Continental Margins; Springer International Publishing: Cham, Switzerland, 2022; pp. 437–447. [Google Scholar]
  169. Bodur, M.N.; Ergin, M. Geochemical Characteristics of the Recent Sediments from the Sea of Marmara. Chem. Geol. 1994, 115, 73–101. [Google Scholar] [CrossRef]
  170. Bourry, C.; Chazallon, B.; Charlou, J.L.; Pierre Donval, J.; Ruffine, L.; Henry, P.; Geli, L.; Çagatay, M.N.; İnan, S.; Moreau, M. Free Gas and Gas Hydrates from the Sea of Marmara, Turkey. Chem. Geol. 2009, 264, 197–206. [Google Scholar] [CrossRef]
  171. Sarıtaş, H.; Çifçi, G.; Géli, L.; Thomas, Y.; Marsset, B.; Henry, P.; Grall, C.; Rochat, A. Gas Occurrence and Shallow Conduit Systems in the Western Sea of Marmara: A Review and New Acoustic Evidence. Geo-Mar. Lett. 2018, 38, 385–402. [Google Scholar] [CrossRef]
  172. Lu, Z.; Zhu, Y.; Zhang, Y.; Wen, H.; Li, Y.; Liu, C. Gas Hydrate Occurrences in the Qilian Mountain Permafrost, Qinghai Province, China. Cold Reg. Sci. Technol. 2011, 66, 93–104. [Google Scholar] [CrossRef]
  173. Zhao, J.; Yu, T.; Song, Y.; Liu, D.; Liu, W.; Liu, Y.; Yang, M.; Ruan, X.; Li, Y. Numerical Simulation of Gas Production from Hydrate Deposits Using a Single Vertical Well by Depressurization in the Qilian Mountain Permafrost, Qinghai-Tibet Plateau, China. Energy 2013, 52, 308–319. [Google Scholar] [CrossRef]
  174. Wang, P.; Zhu, Y.; Lu, Z.; Huang, X.; Pang, S.; Zhang, S. Gas Hydrate Stability Zone Migration Occurred in the Qilian Mountain Permafrost, Qinghai, Northwest China: Evidences from Pyrite Morphology and Pyrite Sulfur Isotope. Cold Reg. Sci. Technol. 2014, 98, 8–17. [Google Scholar] [CrossRef]
  175. Collett, T.S.; Ginsburg, G.D. Gas Hydrates in the Messoyakha Gas Field of the West Siberian Basin—A Re-Examination of the Geologic Evidence. Int. J. Offshore Polar Eng. 1998, 8, ISOPE-98-08-1-022. [Google Scholar]
  176. Uchida, T.; Dallimore, S.; Mikami, J. Occurrences of Natural Gas Hydrates beneath the Permafrost Zone in Mackenzie Delta: Visual and X-ray CT Imagery. Ann. N. Y. Acad. Sci. 2000, 912, 1021–1033. [Google Scholar] [CrossRef]
  177. Dai, S.; Lee, C.; Carlos Santamarina, J. Formation History and Physical Properties of Sediments from the Mount Elbert Gas Hydrate Stratigraphic Test Well, Alaska North Slope. Mar. Pet. Geol. 2011, 28, 427–438. [Google Scholar] [CrossRef]
  178. Winters, W.; Walker, M.; Hunter, R.; Collett, T.; Boswell, R.; Rose, K.; Waite, W.; Torres, M.; Patil, S.; Dandekar, A. Physical Properties of Sediment from the Mount Elbert Gas Hydrate Stratigraphic Test Well, Alaska North Slope. Mar. Pet. Geol. 2011, 28, 361–380. [Google Scholar] [CrossRef]
  179. Onwukwe, S.I.; Duru, U.I. Prospect of Harnessing Associated Gas through Natural Gas Hydrate (NGH) Technology in Nigeria. J. Pet. Gas Eng. 2015, 6, 38–44. [Google Scholar] [CrossRef]
  180. Ganguly, N.; Spence, G.D.; Chapman, N.R.; Hyndman, R.D. Heat Flow Variations from Bottom Simulating Reflectors on the Cascadia Margin. Mar. Geol. 2000, 164, 53–68. [Google Scholar] [CrossRef]
  181. Haacke, R.R.; Westbrook, G.K.; Hyndman, R.D. Gas Hydrate, Fluid Flow and Free Gas: Formation of the Bottom-Simulating Reflector. Earth Planet. Sci. Lett. 2007, 261, 407–420. [Google Scholar] [CrossRef]
  182. Klitzke, P.; Luzi-Helbing, M.; Schicks, J.M.; Cacace, M.; Jacquey, A.B.; Sippel, J.; Scheck-Wenderoth, M.; Faleide, J.I. Gas Hydrate Stability Zone of the Barents Sea and Kara Sea Region. Energy Procedia 2016, 97, 302–309. [Google Scholar] [CrossRef]
  183. Villinger, H.W.; Tréhu, A.M.; Grevemeyer, I. 18. Seafloor Marine Heat Flux Measurements and Estimation of Heat Flux from Seismic Observations of Bottom Simulating Reflectors. In Geophysical Characterization of Gas Hydrates; Society of Exploration Geophysicists: Houston, TX, USA, 2010; pp. 279–300. [Google Scholar]
  184. Kumar, A.; Cook, A.E.; Portnov, A.; Skopec, S.; Frye, M.; Palmes, S. Bottom Simulating Reflections across the Northern Gulf of Mexico Slope. Mar. Pet. Geol. 2024, 165, 106870. [Google Scholar] [CrossRef]
  185. Xu, W.; Ruppel, C. Predicting the Occurrence, Distribution, and Evolution of Methane Gas Hydrate in Porous Marine Sediments. J. Geophys. Res. Solid Earth 1999, 104, 5081–5095. [Google Scholar] [CrossRef]
  186. Holder, G.D. Clathrate Hydrates of Natural Gases, 2nd Ed. J. Am. Chem. Soc. 1998, 120, 11212. [Google Scholar] [CrossRef]
  187. Song, Y.; Lei, Y.; Zhang, L.; Cheng, M.; Miao, L.; Li, C.; Liu, N. Spatial-Temporal Variations of the Gas Hydrate Stability Zone and Hydrate Accumulation Models in the Dongsha Region, China. Front. Mar. Sci. 2022, 9, 982814. [Google Scholar] [CrossRef]
  188. Cai, W.; Guan, W.; Zhan, L.; Lu, H.; Xu, C. Coexistence of Methane Hydrate and Gas Associated with the Difference in Thermodynamic Stabilities of Hydrate in the Pores of Sediments from Shenhu, South China Sea. Energy Fuels 2024, 38, 8612–8619. [Google Scholar] [CrossRef]
  189. Holbrook, W.S.; Hoskins, H.; Wood, W.T.; Stephen, R.A.; Lizarralde, D. Methane Hydrate and Free Gas on the Blake Ridge from Vertical Seismic Profiling. Science 1996, 273, 1840–1843. [Google Scholar] [CrossRef]
  190. Kumar, S. Internal Structure of Hydrated Sediments and BSR as Seismic Evidence of Free Gas Trapping by Hydrate Zone. In Proceedings of the All Days; OTC: Springfield, MO, USA, 1998. [Google Scholar]
  191. Yuan, H.; Wang, Y.; Wang, X. Seismic Methods for Exploration and Exploitation of Gas Hydrate. J. Earth Sci. 2021, 32, 839–849. [Google Scholar] [CrossRef]
  192. Davies, R.J.; Cartwright, J. A Fossilized Opal A to Opal C/T Transformation on the Northeast Atlantic Margin: Support for a Significantly Elevated Palaeogeothermal Gradient during the Neogene? Basin Res. 2002, 14, 467–486. [Google Scholar] [CrossRef]
  193. Gong, Y.; Yang, S.; Liang, J.; Tian, D.; Lu, J.; Deng, W.; Meng, M. Identification of Mass Transport Deposits and Insights into Gas Hydrate Accumulation in the Qiongdongnan Sea Area, Northern South China Sea. J. Mar. Sci. Eng. 2024, 12, 855. [Google Scholar] [CrossRef]
  194. Portnov, A.; Santra, M.; Cook, A.E.; Sawyer, D.E. The Jackalope Gas Hydrate System in the Northeastern Gulf of Mexico. Mar. Pet. Geol. 2020, 111, 261–278. [Google Scholar] [CrossRef]
  195. Baba, K.; Yamada, Y. BSRs and Associated Reflections as an Indicator of Gas Hydrate and Free Gas Accumulation: An Example of Accretionary Prism and Forearc Basin System along the Nankai Trough, off Central Japan. Resour. Geol. 2004, 54, 11–24. [Google Scholar] [CrossRef]
  196. Hart, P.E.; Pohlman, J.W.; Lorenson, T.D.; Edwards, B.D. Beaufort Sea Deep-Water Gas Hydrate Recovery from a Seafloor Mound in a Region of Widespread BSR Occurrence. In Proceedings of the 7th International Conference on Gas Hydrates (ICGH 2011), Edinburgh, Scotland, 17–21 July 2011. [Google Scholar]
  197. Dirgantara, F.; Susilohadi, S.; Iskandar, M.A.S.; Sintuwardhana, Y.A. Methane Hydrate Systems off South Makassar Basin, Indonesia. Geosci. Lett. 2025, 12, 5. [Google Scholar] [CrossRef]
  198. Wei, L.; Cook, A.; You, K. Methane Migration Mechanisms for the Green Canyon Block 955 Gas Hydrate Reservoir, Northern Gulf of Mexico. Am. Assoc. Pet. Geol. Bull. 2022, 106, 1005–1023. [Google Scholar] [CrossRef]
  199. Flemings, P.B.; Phillips, S.C.; Boswell, R.; Collett, T.S.; Cook, A.E.; Dong, T.; Frye, M.; Goldberg, D.S.; Guerin, G.; Holland, M.E.; et al. Pressure Coring a Gulf of Mexico Deep-Water Turbidite Gas Hydrate Reservoir: Initial Results from The University of Texas–Gulf of Mexico 2-1 (UT-GOM2-1) Hydrate Pressure Coring Expedition. Am. Assoc. Pet. Geol. Bull. 2020, 104, 1847–1876. [Google Scholar] [CrossRef]
  200. Riedel, M.; Collett, T.S.; Kumar, P.; Sathe, A.V.; Cook, A. Seismic Imaging of a Fractured Gas Hydrate System in the Krishna–Godavari Basin Offshore India. Mar. Pet. Geol. 2010, 27, 1476–1493. [Google Scholar] [CrossRef]
  201. Pecher, I.A.; Henrys, S.A.; Wood, W.T.; Kukowski, N.; Crutchley, G.J.; Fohrmann, M.; Kilner, J.; Senger, K.; Gorman, A.R.; Coffin, R.B.; et al. Focussed Fluid Flow on the Hikurangi Margin, New Zealand—Evidence from Possible Local Upwarping of the Base of Gas Hydrate Stability. Mar. Geol. 2010, 272, 99–113. [Google Scholar] [CrossRef]
  202. Malinverno, A.; Goldberg, D.S. Testing Short-Range Migration of Microbial Methane as a Hydrate Formation Mechanism: Results from Andaman Sea and Kumano Basin Drill Sites and Global Implications. Earth Planet. Sci. Lett. 2015, 422, 105–114. [Google Scholar] [CrossRef]
  203. Boswell, R.; Collett, T.S.; Frye, M.; Shedd, W.; McConnell, D.R.; Shelander, D. Subsurface Gas Hydrates in the Northern Gulf of Mexico. Mar. Pet. Geol. 2012, 34, 4–30. [Google Scholar] [CrossRef]
  204. Navalpakam, R.S.; Pecher, I.A.; Stern, T. Weak and Segmented Bottom Simulating Reflections on the Hikurangi Margin, New Zealand—Implications for Gas Hydrate Reservoir Rocks. J. Pet. Sci. Eng. 2012, 88–89, 29–40. [Google Scholar] [CrossRef]
  205. Neurauter, T.W.; Bryant, W.R. Seismic Expression of Sedimentary Volcanism on the Continental Slope, Northern Gulf of Mexico. Geo-Mar. Lett. 1990, 10, 225–231. [Google Scholar] [CrossRef]
  206. Terzariol, M.; Park, J.; Castro, G.M.; Santamarina, J.C. Methane Hydrate-Bearing Sediments: Pore Habit and Implications. Mar. Pet. Geol. 2020, 116, 104302. [Google Scholar] [CrossRef]
  207. Crutchley, G.J.; Hillman, J.I.T.; Kroeger, K.F.; Watson, S.J.; Turco, F.; Mountjoy, J.J.; Davy, B.; Woelz, S. Both Longitudinal and Transverse Extension Controlling Gas Migration Through Submarine Anticlinal Ridges, New Zealand’s Southern Hikurangi Margin. J. Geophys. Res. Solid Earth 2023, 128, e2022JB026279. [Google Scholar] [CrossRef]
  208. Posewang, J.; Mienert, J. High-Resolution Seismic Studies of Gas Hydrates West of Svalbard. Geo-Mar. Lett. 1999, 19, 150–156. [Google Scholar] [CrossRef]
  209. Andressen, K.; Mienert, J.; Bryn, P.; Singh, S.C. A Double Gas-Hydrate Related Bottom Simulating Reflector at the Norwegian Continental Margin. Ann. N. Y. Acad. Sci. 2000, 912, 126–135. [Google Scholar] [CrossRef]
  210. Wu, S.; Zhang, G.; Huang, Y.; Liang, J.; Wong, H.K. Gas Hydrate Occurrence on the Continental Slope of the Northern South China Sea. Mar. Pet. Geol. 2005, 22, 403–412. [Google Scholar] [CrossRef]
  211. Auguy, C.; Calvès, G.; Calderon, Y.; Brusset, S. Seismic Evidence of Gas Hydrates, Multiple BSRs and Fluid Flow Offshore Tumbes Basin, Peru. Mar. Geophys. Res. 2017, 38, 409–423. [Google Scholar] [CrossRef]
  212. Han, S.; Bangs, N.L.; Hornbach, M.J.; Pecher, I.A.; Tobin, H.J.; Silver, E.A. The Many Double BSRs across the Northern Hikurangi Margin and Their Implications for Subduction Processes. Earth Planet. Sci. Lett. 2021, 558, 116743. [Google Scholar] [CrossRef]
  213. Li, L.; Liu, H.; Zhang, X.; Lei, X.; Sha, Z. BSRs, Estimated Heat Flow, Hydrate-Related Gas Volume and Their Implications for Methane Seepage and Gas Hydrate in the Dongsha Region, Northern South China Sea. Mar. Pet. Geol. 2015, 67, 785–794. [Google Scholar] [CrossRef]
  214. Qian, J.; Wang, X.; Collett, T.S.; Guo, Y.; Kang, D.; Jin, J. Downhole Log Evidence for the Coexistence of Structure II Gas Hydrate and Free Gas below the Bottom Simulating Reflector in the South China Sea. Mar. Pet. Geol. 2018, 98, 662–674. [Google Scholar] [CrossRef]
  215. Jin, J.; Wang, X.; He, M.; Li, J.; Yan, C.; Li, Y.; Zhou, J.; Qian, J. Downward Shift of Gas Hydrate Stability Zone Due to Seafloor Erosion in the Eastern Dongsha Island, South China Sea. J. Oceanol. Limnol. 2020, 38, 1188–1200. [Google Scholar] [CrossRef]
  216. Davies, R.J.; Morales Maqueda, M.Á.; Li, A.; Ganopolski, A. Millennial-Scale Shifts in the Methane Hydrate Stability Zone Due to Quaternary Climate Change. Geology 2017, 45, 1027–1030. [Google Scholar] [CrossRef]
  217. Chen, J.; Zhao, W.; Tong, S.; Azevedo, L.; Wu, N.; Liu, B.; Xu, H.; Gong, J.; Liao, J.; Liang, J.; et al. Study on the Occurrence of Double Bottom Simulating Reflectors in the Makran Accretionary Zone. J. Mar. Sci. Eng. 2025, 13, 68. [Google Scholar] [CrossRef]
  218. Tinivella, U.; Giustiniani, M. Variations in BSR Depth Due to Gas Hydrate Stability versus Pore Pressure. Glob. Planet. Change 2013, 100, 119–128. [Google Scholar] [CrossRef]
  219. Riedel, M.; Collett, T.S. Observed Correlation between the Depth to Base and Top of Gas Hydrate Occurrence from Review of Global Drilling Data. Geochem. Geophys. Geosyst. 2017, 18, 2543–2561. [Google Scholar] [CrossRef]
  220. Foucher, J.-P.; Nouzé, H.; Henry, P. Observation and Tentative Interpretation of a Double BSR on the Nankai Slope. Mar. Geol. 2002, 187, 161–175. [Google Scholar] [CrossRef]
  221. Zhang, W.; Liang, J.; Qiu, H.; Deng, W.; Meng, M.; He, Y.; Huang, W.; Liang, J.; Lin, L.; Wang, L.; et al. Double Bottom Simulating Reflectors and Tentative Interpretation with Implications for the Dynamic Accumulation of Gas Hydrates in the Northern Slope of the Qiongdongnan Basin, South China Sea. J. Asian Earth Sci. 2022, 229, 105151. [Google Scholar] [CrossRef]
  222. Gupta, S.; Deusner, C.; Burwicz-Galerne, E.; Haeckel, M. Numerical Analysis of the Dynamic Gas Hydrate System and Multiple BSRs in the Danube Paleo-Delta, Black Sea. Mar. Geol. 2024, 469, 107221. [Google Scholar] [CrossRef]
  223. Wang, X.; Wang, W.; Jin, J.; Zhou, J.; Kuang, Z.; Hu, G.; Zhang, Z.; Li, S. Features and Controlling Factors on the Dynamic Adjustment of Marine Gas Hydrate System. Acta Geol. Sin. 2024, 98, 2541–2556. [Google Scholar] [CrossRef]
  224. Paganoni, M.; Cartwright, J.A.; Foschi, M.; Shipp, R.C.; Van Rensbergen, P. Structure II Gas Hydrates Found below the Bottom-simulating Reflector. Geophys. Res. Lett. 2016, 43, 5696–5706. [Google Scholar] [CrossRef]
  225. Goto, S.; Matsubayashi, O.; Nagakubo, S. Simulation of Gas Hydrate Dissociation Caused by Repeated Tectonic Uplift Events. J. Geophys. Res. Solid Earth 2016, 121, 3200–3219. [Google Scholar] [CrossRef]
  226. Popescu, I.; De Batist, M.; Lericolais, G.; Nouzé, H.; Poort, J.; Panin, N.; Versteeg, W.; Gillet, H. Multiple Bottom-Simulating Reflections in the Black Sea: Potential Proxies of Past Climate Conditions. Mar. Geol. 2006, 227, 163–176. [Google Scholar] [CrossRef]
  227. Shaddox, H.R.; Schwartz, S.Y. Subducted Seamount Diverts Shallow Slow Slip to the Forearc of the Northern Hikurangi Subduction Zone, New Zealand. Geology 2019, 47, 415–418. [Google Scholar] [CrossRef]
  228. Nakajima, J.; Uchida, N. Repeated Drainage from Megathrusts during Episodic Slow Slip. Nat. Geosci. 2018, 11, 351–356. [Google Scholar] [CrossRef]
  229. Li, C.; Zhan, L.; Lu, H. Mechanisms for Overpressure Development in Marine Sediments. J. Mar. Sci. Eng. 2022, 10, 490. [Google Scholar] [CrossRef]
  230. Zhang, W.; Liang, J.; Wan, Z.; Su, P.; Huang, W.; Wang, L.; Lin, L. Dynamic Accumulation of Gas Hydrates Associated with the Channel-Levee System in the Shenhu Area, Northern South China Sea. Mar. Pet. Geol. 2020, 117, 104354. [Google Scholar] [CrossRef]
  231. Musgrave, R.J.; Bangs, N.L.; Larrasoaña, J.C.; Gràcia, E.; Hollamby, J.A.; Vega, M.E. Rise of the Base of the Gas Hydrate Zone since the Last Glacial Recorded by Rock Magnetism. Geology 2006, 34, 117. [Google Scholar] [CrossRef]
Figure 2. Global distribution and morphological types of BSR-developed regions. Note: the five-lobed star-shaped symbol (*) indicates double/multiple BSRs, with its color representing the morphological features of BSR1.
Figure 2. Global distribution and morphological types of BSR-developed regions. Note: the five-lobed star-shaped symbol (*) indicates double/multiple BSRs, with its color representing the morphological features of BSR1.
Jmse 13 01137 g002
Figure 3. Correlation analysis between BSR depths from global BSR field observations and predicted BGHSZ values from the CSMHYD program: (a) Observations vs. phase boundaries; (b) ΔZ distribution; and (c,d) Correlations under saltwater and pure water conditions.
Figure 3. Correlation analysis between BSR depths from global BSR field observations and predicted BGHSZ values from the CSMHYD program: (a) Observations vs. phase boundaries; (b) ΔZ distribution; and (c,d) Correlations under saltwater and pure water conditions.
Jmse 13 01137 g003
Figure 4. Morphological types of BSRs and their corresponding geological formation models. The left column (a1e1) shows seismic reflection profiles from various study areas, and the right column (a2e2) illustrates corresponding conceptual geological models: (a1,d1) Gulf of Mexico, modified from Kumar et al. [184] and Portnov et al. [194]; (b1,e1) Colombian Caribbean Margin, modified from Leslie et al. [76]; and (c1) Sackville Spur, Offshore Newfoundland, modified from Mosher et al. [94]. In the seismic sections, white arrows indicate the locations of BSRs.
Figure 4. Morphological types of BSRs and their corresponding geological formation models. The left column (a1e1) shows seismic reflection profiles from various study areas, and the right column (a2e2) illustrates corresponding conceptual geological models: (a1,d1) Gulf of Mexico, modified from Kumar et al. [184] and Portnov et al. [194]; (b1,e1) Colombian Caribbean Margin, modified from Leslie et al. [76]; and (c1) Sackville Spur, Offshore Newfoundland, modified from Mosher et al. [94]. In the seismic sections, white arrows indicate the locations of BSRs.
Jmse 13 01137 g004
Figure 5. Types of the BSR in structural settings and their corresponding mechanisms: (a) Ridge-type; (b) Buried anticline-type; and (c) Accretionary prism-type.
Figure 5. Types of the BSR in structural settings and their corresponding mechanisms: (a) Ridge-type; (b) Buried anticline-type; and (c) Accretionary prism-type.
Jmse 13 01137 g005
Figure 6. Sediment grain-size compositions from global BSR observations are plotted on a ternary diagram (sand–silt–clay). Numerical labels in the diagram correspond to locations in Figure 2 and Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8, while the color and shape of the markers indicate BSR morphology and its geological setting.
Figure 6. Sediment grain-size compositions from global BSR observations are plotted on a ternary diagram (sand–silt–clay). Numerical labels in the diagram correspond to locations in Figure 2 and Table 2, Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8, while the color and shape of the markers indicate BSR morphology and its geological setting.
Jmse 13 01137 g006
Figure 7. Mechanism of BGHSZ depth variation induced by rapid sedimentation. The brown curve marks the gas hydrate phase boundary. The blue curve indicates water temperature above the seafloor, while the blue line represents sub-bottom temperature (dashed: pre-sedimentation; solid: post-sedimentation). Dark and light green dashed lines denote post-depositional (Seafloor1) and pre-depositional (Seafloor2) seafloor depths. BGHSZ1 and BGHSZ2, defined by intersections of the phase boundary with post- and pre-sedimentation temperature lines, respectively, delineate the BGHSZ. This schematic highlights how rapid deposition modifies seafloor temperature, inducing an upward migration of the BGHSZ.
Figure 7. Mechanism of BGHSZ depth variation induced by rapid sedimentation. The brown curve marks the gas hydrate phase boundary. The blue curve indicates water temperature above the seafloor, while the blue line represents sub-bottom temperature (dashed: pre-sedimentation; solid: post-sedimentation). Dark and light green dashed lines denote post-depositional (Seafloor1) and pre-depositional (Seafloor2) seafloor depths. BGHSZ1 and BGHSZ2, defined by intersections of the phase boundary with post- and pre-sedimentation temperature lines, respectively, delineate the BGHSZ. This schematic highlights how rapid deposition modifies seafloor temperature, inducing an upward migration of the BGHSZ.
Jmse 13 01137 g007
Figure 8. Different genetic mechanisms of double/multiple BSRs: (a) Black Sea, modified from Zander et al. [165]; (b) Nankai Trough, modified from Wang et al. [223]; (c) Shenhu Area, South China Sea, modified from Jin et al. [40]; (d) Hikurangi Margin, modified from Han et al. [212]; (e) Qiongdongnan Basin, South China Sea, modified from Zhang et al. [221]; (f) NW Borneo, modified from Paganoni et al. [224].
Figure 8. Different genetic mechanisms of double/multiple BSRs: (a) Black Sea, modified from Zander et al. [165]; (b) Nankai Trough, modified from Wang et al. [223]; (c) Shenhu Area, South China Sea, modified from Jin et al. [40]; (d) Hikurangi Margin, modified from Han et al. [212]; (e) Qiongdongnan Basin, South China Sea, modified from Zhang et al. [221]; (f) NW Borneo, modified from Paganoni et al. [224].
Jmse 13 01137 g008
Table 1. Seismic identification of BSR in gas hydrate systems.
Table 1. Seismic identification of BSR in gas hydrate systems.
CategoryKey Features/ParametersDescriptionStrengthsLimitationsRefs.
Conventional seismic-section analysisMorphological Features- Thermal conduction control: BSR is sub-parallel to the seafloor/isotherms.
- Thermal convection control: BSR appears as an upward-convex structure and often develops in fluid migration zones.
- Usually a cross-cutting, high-amplitude reflector (>30% of seafloor-reflection amplitude).
Visually straightforward; enables rapid delineation of hydrate-bearing provinces over large areas.- Highly dependent on data quality; low signal-to-noise ratios or low dominant frequencies can render the BSR hard to recognize
- In structurally complex settings the BSR may be masked or confused with other stratigraphic boundaries.
- Where free gas is limited, the reflector weakens or disappears, leading to missed detections.
[25,51]
Polarity FeaturesA hydrate-related BSR typically shows negative polarity (central negative lobe flanked by positive side lobes), distinguishing it from diagenetic BSRs with positive polarity.Directly reflects the impedance reversal at the hydrate/gas interface; a useful auxiliary discriminator.- May be confused with negative-polarity reflections generated solely by underlying high-saturation gas layers or overlaying carbonates.
Associated reflection phenomenaPositive-polarity reflection below the BSRResulting from significant impedance contrasts between free gas zones and underlying sediments.To enhance the reliability of hydrate-free gas system identification by incorporating multiple lines of evidence.- Characterized by weak amplitudes and simple geometries, these features are observable only in high-resolution, low-noise profiles.
- They can be easily mistaken for other bright spots or stratigraphic boundaries; therefore, meticulous validation is imperative during interpretation.
[33,34]
Positive-polarity reflection at the hydrate top boundaryAssociated with impedance contrasts between hydrate-bearing sediments and overlying normal sediments.
Acoustic blanking zone below the BSRAttributed to seismic attenuation linked to high gas concentrations within gas chimneys or mud diapirs.
Acoustic blanking zone within the hydrate-bearing layerDue to the homogeneous nature of hydrate-bearing sediments, leading to reduced seismic reflection amplitudes.
Seismic-attribute analysisInstantaneous Amplitude- Highlights prominent BSR reflections and amplitude anomalies within hydrate-bearing sediments.
- Independent of seismic phase and thus polarity-insensitive.
To overcome the shortcomings of traditional methods and facilitate the detection of the weak or discontinuous BSR.- High data quality requirements; noise or multiples can introduce artifacts.
- Subjectivity in the selection of attribute parameters.
- AVO analysis requires well-sampled multi-offset data and is unsuitable for datasets with limited offset coverage or conventional single-offset data.
[43]
Instantaneous Phase- Improves BSR visibility when cross-cutting strata.
- Limited sensitivity to parallel bedding.
[44,45]
Instantaneous Frequency- Free gas absorption leads to low-frequency shadowing, aiding in free gas identification.[36]
AVO Attributes- Intercept parameter facilitates BSR identification.
- Fluid factor is highly sensitive to free gas detection.
[50]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shi, S.; Zhan, L.; Cai, W.; Yang, R.; Lu, H. Bottom-Simulating Reflectors (BSRs) in Gas Hydrate Systems: A Comprehensive Review. J. Mar. Sci. Eng. 2025, 13, 1137. https://doi.org/10.3390/jmse13061137

AMA Style

Shi S, Zhan L, Cai W, Yang R, Lu H. Bottom-Simulating Reflectors (BSRs) in Gas Hydrate Systems: A Comprehensive Review. Journal of Marine Science and Engineering. 2025; 13(6):1137. https://doi.org/10.3390/jmse13061137

Chicago/Turabian Style

Shi, Shiyuan, Linsen Zhan, Wenjiu Cai, Ran Yang, and Hailong Lu. 2025. "Bottom-Simulating Reflectors (BSRs) in Gas Hydrate Systems: A Comprehensive Review" Journal of Marine Science and Engineering 13, no. 6: 1137. https://doi.org/10.3390/jmse13061137

APA Style

Shi, S., Zhan, L., Cai, W., Yang, R., & Lu, H. (2025). Bottom-Simulating Reflectors (BSRs) in Gas Hydrate Systems: A Comprehensive Review. Journal of Marine Science and Engineering, 13(6), 1137. https://doi.org/10.3390/jmse13061137

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop