Skip to main content

ORIGINAL RESEARCH article

Front. Earth Sci., 04 January 2023
Sec. Solid Earth Geophysics
This article is part of the Research Topic Rock Physics of Unconventional Reservoirs View all 11 articles

P-wave anelasticity in hydrate-bearing sediments based on a triple-porosity model

  • 1School of Earth Sciences and Engineering, Hohai University, Nanjing, China
  • 2National Institute of Oceanography and Applied Geophysics—OGS, Trieste, Italy

P-wave anelasticity (attenuation and dispersion) of hydrate-bearing sediments depends on several factors, namely the properties of the mineral components, hydrate content and morphology, and fluid saturation. Anelasticity is analyzed with a triple-porosity model (stiff pores, clay micropores and hydrate micropores), by considering hydrate as an additional solid skeleton. We relate the hydrate volume ratio, porosity and radii of the hydrate inclusion and clay mineral to the P-wave velocity and attenuation. The model takes wave-induced local fluid flow (mesoscopic loss) at the grain contacts into account. The results are compared with those of a double-porosity and load-bearing models, and verified with well-log data from Offshore Drilling Program sites 1247B and 1250F, and data reported in Nankai Trough, Japan. Model results and data show a good agreement.

Introduction

Gas hydrate is an ice-like crystalline medium with a microporous structure composed of gas and water molecules that are formed at low temperature, high pressure and certain gas saturation (Sloan, 1990). Identification of hydrate reservoirs in engineering applications worldwide mainly relies on seismic exploration techniques. The existence of hydrates highly affects the acoustic wave velocity and attenuation. Generally, hydrate-bearing sediments show high compressional (P-) and shear (S-) wave velocities. These velocities and attenuation are usually adopted to estimate the presence of hydrates (Waite et al., 2009). With the increasing hydrate content, the wave velocity increases. Moreover, the morphology and distribution of hydrate also have an effect on the velocities (Ecker et al., 1998; Ecker et al., 2000). However, the relation between attenuation and hydrate content is more complex, as it is associated with mechanisms due to different microporous hydrate forms (Best et al., 2013).

Rock physics is an effective approach to describe the quantitative relation between the rock microstructure and the wave properties. A suitable model associated with the hydrate morphology could be helpful to improve the accuracy in the estimation of hydrate content (Pan et al., 2019). In fact, the hydrate-bearing sediment is a three-phase porous medium, composed of a rock frame saturated with a fluid (usually water) and hydrates (Liu et al., 2021). Biot (1962) considered wave propagation in fully saturated porous media including anisotropy and viscoelasticity, and a loss mechanism related to the differential motion between the frame and the fluid. Stoll and Bryan. (2009) was the first to systematically apply Biot’s theory to marine sediments. Carcione and Gei (2004) proposed a Biot-type theory of two solids and one fluid, in which hydrate is considered as a second skeleton (frame) and water is the pore fluid—grain cementation and friction between the two frames have been considered (see also, Carcione et al. (2005); Gei and Carcione (2003); Gei et al. (2022)). This model has also been used by Guerin and Goldberg (2005).

On the other hand, Ba et al. (2011); Ba et al. (2016) proposed a double-porosity model to describe attenuation due to local fluid flow between soft and stiff pores (mesoscopic or microscopic loss). For the same purpose, Zhang et al. (2017) presented an alternative model based on the triple-porosity structure of sand, gravel and mudstone while, Zhang et al. (2016) applied the BISQ (Biot/squirt) model specifically to marine unconsolidated hydrate-bearing sediments. They found that wave velocity and attenuation increase with increasing hydrate content, in agreement with some measurement data (Chand and Minshull, 2004; Guerin and Goldberg, 2005; Matsushima, 2006), and that porosity has a weak effect on attenuation. Zhan et al. (2022) discussed and compared the feasibility and limitations of existing rock physics models. The combination of different models may be more conducive to explaining the mechanism of attenuation.

Attenuation also depends on the hydrate morphology and microstructure (Priest et al., 2009), as shown by laboratory measurements, including local viscous fluid flow related to the microporous structure of hydrate containing gas and water (Best et al., 2013). Leurer and Brown (2008) proposed a model to explain the viscoelasticity generated by local fluid flow in the presence of clay at grain contacts. Marín-Moreno et al. (2017) developed the hydrate-bearing effective sediment model (HBES) to analyze various loss mechanisms, including those caused by squirt flow in microporous hydrate, viscoelasticity of the hydrate frame and Biot global flow. They analyzed the effect of hydrate morphology on attenuation by comparing results between sediments with and without hydrates. Sahoo et al. (2019) performed high-precision ultrasonic pulse-echo measurements of wave velocity and attenuation in hydrate-bearing sediments. Li et al. (2015) studied the effect of clay content on the mechanical properties of these sediments by applying the tests of multi-stage loading triaxial compression and hydrate decomposition. On the basis of six microscopic hydrate morphologies, Pan et al. (2019) obtained rock-physics templates (RPTs), based on an amplitude-variation with offset (AVO) analysis, and predicted hydrate content, porosity and clay content of permafrost-associated hydrate-bearing sediments at Mount Elbert, North Slope of Alaska.

The hydrate distribution in the pore space is important. In the formation process, gas hydrate is present in different forms due to the influence of the geological setting, formation pressure and geothermal gradient. Ecker et al. (1998) proposed three types of hydrate distributions, namely, grain-contact cementing, grain coating and absent in the grain contacts. Dai et al. (2004) proposed six distributions: grain-contact cementing, grain coating, supporting matrix/grains, pore-filling, matrix and inclusions, and nodules/fracture fillings. Zhan and Matsushima. (2018) considered four distributions: grain-contact cementing, grain coating, load-bearing and pore-filling (Schicks et al., 2006). Three microscopic distribution patterns of hydrates, namely pore filling, contact or encapsulated cementation and load-bearing hydrates, are discussed by Waite et al. (2009).

Understanding the effect of anelasticity on the acoustic properties is not clear, mainly because attenuation behaves differently at different frequency bands. Here, we consider the local fluid flow between stiff pores, hydrates, and clay, based on a triple-porosity model (Zhang et al., 2017). In addition to the double-porosity model, the new model considers the effects of clay micropores. The results agree with log data, providing an effective approach to model the P-wave anelasticity mechanisms of hydrate-bearing sediments.

The model

We consider the main frame or skeleton containing intergranular pores as the host phase, hydrate and clay as two different types of multi-pore inclusions, and describe the hydrate-bearing sediments with a triple-porosity model. It has been observed that hydrate can cement the mineral grain and contributes to the solid skeleton or being part of pore-filling material. Cementation decreases the porosity and increases the bulk modulus of the skeleton. Figure 1 shows three cases where hydrate is 1) part of the pore infill; 2) part of the frame; 3) cementing the grains. We consider the case in panel 2).

FIGURE 1
www.frontiersin.org

FIGURE 1. Scheme showing the three hydrate morphologies.

In this case, the hydrate-bearing sediment can be regarded as a composite of three skeletons: rock (minerals), hydrate and clay. Basically, hydrate and clay reduce the bulk porosity and combine with the minerals as shown in Figure 2, and then the resulting frame is saturated.

FIGURE 2
www.frontiersin.org

FIGURE 2. Diagram showing hydrate as part of the solid frame (skeleton).

As stated above, the understanding on wave-loss mechanisms of hydrate-bearing sediments is still limited (Best et al., 2013). Figure 3 shows the mechanisms considered in our triple-porosity model, i.e., local fluid flow between the rock skeleton and hydrate and clay frames, and the classical Biot global flow (Zhang et al., 2022).

FIGURE 3
www.frontiersin.org

FIGURE 3. Attenuation mechanisms of the triple-porosity model.

Properties of fluid

If hydrate is part of the frame, the pore fluid is a mixture of water and free gas, such that its effective bulk modulus is (Wood, 1955; Wood et al., 2000; Liu et al., 2017)

Kf=SwKw+SgKg1(1)

where Sw and Sg are the water and free gas saturations, Kg and Kw are the respective bulk moduli with Sw+Sg=1.

The effective density of the fluid is

ρf=ρwSw+ρgSg(2)

where ρf, ρw and ρg are the densities of fluid, water and gas, respectively.

Properties of solid phase (composite mineral)

According to the Hill average (Hill, 1952), the moduli of the solid phase considering the presence of hydrate are (Helgerud et al., 1999; Ecker et al., 2000)

Ks=12i=14fiKi+i=14fiKi1(3a)
Gs=12i=14fiGi+i=14fiGi1(3b)

where i=1,2,3and4 indicate calcite, quartz, hydrate and clay, respectively, Ki and Gi are the respective bulk and shear moduli of the ith constituent, and fi is the volume fraction of the ith constituent, with 14fi=1.

Properties of the frame

Let us define the volume ratios, local porosities and absolute porosities of the host phase made of quartz and calcite, hydrate skeleton and clay skeleton as ν2,ν1 and ν3, ϕ20, ϕ10 and ϕ30, and ϕ2, ϕ1 and ϕ, respectively, with ϕ1=ϕ10ν1, ϕ2=ϕν2 and ϕ3=ϕ30ν3 (Zhang et al., 2017; Wang et al., 2021; Zhang et al., 2021), and

ν2=1ν1ν3(4a)
f3ν11ϕ10=f4ν31ϕ30=f1+f2ν21ϕ20(4b)

The porosity of the host phase is

ϕ2=ϕϕ1ϕ3(5)

where ϕ is the porosity of the rock with hydrate formation.

Compared to the Hashin-Shtrikman upper bound, the Hashin-Shtrikman lower bound is appropriate for estimating the elastic moduli of submarine sediments, where the soft components (clay or soft minerals) are majorly distributed surrounding the stiff grains. The dry-rock elastic moduli are obtained by the modified Hashin-Shtrikman lower bound (Ecker et al., 1998; Dvorkin et al., 1999; Helgerud et al., 1999)

Kb=ϕ/ϕcKHM+43GHM+1ϕ/ϕcKs+43GHM143GHMϕ<ϕc(6a)
Kb=1ϕ/1ϕcKHM+43GHM+ϕϕc/1ϕc43GHM143GHMϕ>ϕc(6b)
Gb=ϕ/ϕcGHM+Z+1ϕ/ϕcGs+Z1Zϕ<ϕc(7a)
Gb=1ϕ/1ϕcGHM+Z+ϕϕc/1ϕcZ1Zϕ>ϕc(7b)

where KHM and GHM are the bulk and shear moduli of the rock under the critical porosity, respectively, and

Z=GHM69KHM+8GHMKHM+2GHM(8a)
KHM=Gs2n21ϕc218π21σ2P13(8b)
GHM=54σ52σ3Gs2n21ϕc22π21σ2P13(8c)

where ϕc is the critical porosity, ranging from 0.36 to 0.4, n is the coordination number (the average number of contacts per grain, ranging from 8 to 9.5), P is the effective stress, P=1ϕρsρfgh, ρs is the average density of the skeleton, ρs=1mfiρi, where ρi is the density of the ith constituent, h is the depth below sea floor, g is the acceleration of gravity, and σ is the Poisson ratio of the solid phase.

Properties of the saturated sediment

There are two approaches to relate the hydrate content to the P-wave velocity. One method is the use of empirical relations, such as the time-average equation (Wyllie et al., 1958) and Lee weighted equation (Lee et al., 1996; Lee and Collett, 2009), by combining the Wood and time-average equations (Helgerud et al., 1999). Other approaches are poroelasticity and effective medium theories (Helgerud et al., 1999). These methods involve input parameters which are difficult to be obtained in actual applications (Hu et al., 2010).

The double-porosity theory considers that in the process of seismic wave propagation, the micropore structure of the hydrate phase induces a local flow between these micropores and the stiff pores and causes energy loss and velocity dispersion (see Figure 4). This theory, however, ignores the presence of clay, which may also cause flow (Wang et al., 2021). The inclusion of clay leads to the triple-porosity model (see red arrow in Figure 4).

FIGURE 4
www.frontiersin.org

FIGURE 4. Schematic diagram of the proposed triple-porosity model.

Zhang et al. (2017) developed a theory to model the properties of a saturated medium, where the local fluid-flow mechanisms, responsible for wave attenuation and dispersion, are considered. Based on Hamilton’s principle, the dynamical equations can be obtained from the strain energy, kinetic energy and dissipation potential of a triple-porosity medium. The differential equations of motion, extended to the case of hydrate-bearing sediments, are

Gb2u+A+Gbe+Q1ξ1+ϕ2ζ12+Q2ξ2ϕ1ζ12+ϕ3ζ23+Q3ξ3ϕ2ζ23=ρ00u¨+ρ01U¨1+ρ02U¨2+ρ03U¨3+b1u˙U˙1+b2u˙U˙2+b3u˙U˙3(9a)
Q1e+R1ξ1+ϕ2ζ12=ρ01u¨+ρ11U¨1b1u˙U˙1(9b)
Q2e+R2ξ2ϕ1ζ12+ϕ3ζ23=ρ02u¨+ρ22U¨2b2u˙U˙2(9c)
Q3e+R3ξ3ϕ2ζ23=ρ03u¨+ρ33U¨3b3u˙U˙3(9d)
13ρfR122ζ¨12ϕ22ϕ115+ϕ10ϕ20+13η5κ1+ηκ2R122ζ˙12ϕ22ϕ1ϕ10=ϕ2Q1e+R1ξ1+ϕ2ζ12ϕ1Q2e+R2ξ2ϕ1ζ12+ϕ3ζ23(9e)
ϕ33ρfR232ζ¨23ϕ2215+ϕ30ϕ20+13ηκ2+η5κ3R232ζ˙23ϕ22ϕ3ϕ30=ϕ3Q2e+R2ξ2ϕ1ζ12+ϕ3ζ23ϕ2Q3e+R3ξ3ϕ2ζ23(9f)

where u˙, U˙1, U˙2, and U˙3 are the displacement vector of the frame and the average fluid displacement vectors in the hydrate internal pores, intergranular pores and clay micropores, respectively; e, ξ1, ξ2, ξ3 are the displacement divergence fields of the solid and fluids in the three types of pore systems, respectively; ζ12, ζ23 are the bulk strain increments caused by the local flow between the hydrate micropores and intergranular pores, and the local flow between the clay micropores and intergranular pores, respectively; ρ00, ρ01, ρ02, ρ03, ρ11, ρ22, and ρ33 are the Biot density coefficients; b1, b2, and b3 are dissipation coefficients (Biot, 1962; Zhang et al., 2017; see Appendix A); Q1, Q2 and Q3, are the elastic parameters of coupled solid and fluid, A is the elastic parameter of solid phase, and R1, R2, and R3 are the elastic parameters of flow phase (see Appendix A); κ1, κ2 and κ3 denote the permeabilities of the hydrate skeleton, host phase and clay skeleton, respectively; η is fluid viscosity, R12 and R23 denote the hydrate inclusion radius and clay inclusion radius, respectively. The above equations are solved with a plane-wave analysis to obtain the phase velocity and attenuation (see Appendix B) (Ba et al., 2011; Ba et al., 2012; Carcione, 2022).

Example

The minerals are calcite, quartz, clay and hydrate, with volume fractions of f1 = 4%, f2 = 70%, f4 = 20% and f3 = 6%, respectively. The rock porosity with hydrate formation is ϕ = 35%, and the free gas saturation is Sg = 2%. The volume ratios of the hydrate and clay frames are ν1 = 4% and ν3 = 13.1%, respectively, and the corresponding local porosities are ϕ10 = 2% and ϕ30 = 0.5%, respectively. The bulk modulus of hydrate, rock and clay frames are Kb1 = 0.76 GPa, Kb2 = 1.27 GPa, and Kb3 = 1.02 GPa, respectively, the permeability of the host phase is κ2 = 1×10–11 m2, the permeability of the hydrate or clay frames is κ1 = κ3 = 1×10–13 m2, and the fluid viscosity is η = 0.001 kg/(m · s). Table 1 shows the properties of the different phases.

TABLE 1
www.frontiersin.org

TABLE 1. Properties of the phases (Helgerud et al., 1999).

The energy loss caused by the fluid flow depends on the radius of the hydrate inclusions with micropores. Figure 5 shows the results of the double-porosity theory (clay is considered part of the host phase and hydrate is an inclusion), where we can observe a single inflection point and attenuation peak. With increasing radius of these inclusions, the local fluid-flow attenuation peak moves to the low frequencies. The global fluid-flow peak, occurring at high frequencies, is much weaker, almost negligible.

FIGURE 5
www.frontiersin.org

FIGURE 5. Frequency dependence of the P-wave velocity (A) and attenuation (B) corresponding to the double-porosity model with two radii of the hydrate inclusions.

On the other hand, Figure 6 shows the results of the triple-porosity theory, which exhibits the two local fluid-flow mechanisms, between the stiff pores and the soft pores of clay and hydrate phases. The global Biot peak is also present. The clay inclusion radius is R23 = 0.005 cm. The peak due to hydrate merges with the global flow peak when the radius of the hydrate inclusions increase, while the peaks due to clay is not affected.

FIGURE 6
www.frontiersin.org

FIGURE 6. Frequency dependence of the P-wave velocity (A) and attenuation (B) corresponding to the triple-porosity model with two radii of the hydrate inclusions.

Figures 7A, B show the P-wave velocity and dissipation factor as a function of frequency for different clay inclusion radii, respectively. Changes can be observed at high frequencies. When the peaks are close, higher attenuation is observed.

FIGURE 7
www.frontiersin.org

FIGURE 7. Effect of the clay inclusion radius on the P-wave velocity dispersion (A) and attenuation (B). The radius of the hydrate inclusion is R12 = 0.05 cm, and the rock porosity is ϕ = 20%.

Figures 8A, B show the P-wave velocity and dissipation factor as a function of frequency for different hydrate inclusion radii, respectively, where we can see differences at middle frequencies.

FIGURE 8
www.frontiersin.org

FIGURE 8. Effect of the hydrate inclusion radius on the P-wave velocity dispersion (A) and attenuation (B). The radius of the clay inclusion is R23 = 0.005 cm, and the rock porosity is ϕ = 20%.

Figures 9A, B show the P-wave velocity and dissipation factor as a function of frequency for different porosities, respectively, where we can see that increasing porosity enhances the loss due to the local flow related to clay and the global flow.

FIGURE 9
www.frontiersin.org

FIGURE 9. Effect of porosity on the P-wave velocity dispersion (A) and attenuation (B). The radius of the hydrate inclusion is R12 = 0.005 cm, and the radius of the clay inclusion is R23 = 0.05 cm.

Finally, Figures 10A, B show the P-wave phase velocity and dissipation factor as a function of frequency for different hydrate volume ratios, respectively. With the increase of hydrate volume ratio, the P-wave anelasticity due to the hydrate inclusions increase, while those of the clay local flow and global flow decrease. The hydrate volume ratio has no apparent effect on the characteristic frequencies of the peaks.

FIGURE 10
www.frontiersin.org

FIGURE 10. Effect of the hydrate volume ratios on the P-wave velocity dispersion (A) and attenuation (B). The radius of the hydrate inclusions is R12 = 0.005 cm, and the radius of the clay inclusions is R23 = 0.05 cm.

Comparison with well-log data

The Offshore Drilling Program (ODP) drilled through a gas hydrate stabilization zone on the Cascadia edge off Oregon, providing information on the physical properties of hydrate-bearing sediments. The present model is applied to log data of wells 1247B and 1250F of the ODP204 cruise by Pan et al. (2019) and to data obtained by Zhan and Matsushima. (2018) in the Nankai Trough, in Japan.

ODP data

Figures 11A, B show the theoretical and measured (symbols) P-wave velocities as a function of porosity and hydrate saturation, where Sh denotes hydrate saturation (with the relation of ν1=ϕSh ), between 0 and 19%, corresponding to wells 1247B and 1250F, respectively. The variations of scatters with respect to the colorbar reflect the trend that the P-wave velocity of hydrate reservoir rocks decreases with increasing porosity and increases with hydrate saturation. The agreement is good, with the velocity decreasing with increasing porosity and decreasing hydrate saturation.

FIGURE 11
www.frontiersin.org

FIGURE 11. Measured P-wave velocity (symbols) as a function of porosity for different hydrate saturations in wells 1247B (A) and 1250F (B) compared to the model results (solid lines).

Figures 12A, B shows the P-wave velocity as a function of the hydrate saturation in the two wells. In well 1247B the porosity range is ϕ = 0.525–0.535, with an average of 0.53, while that of well 1250F is ϕ = 0.545–0.555 with an average of 0.55. The clay volume ratio in both wells is 0.2 (Pan et al., 2019). Again, the agreement is satisfactory. There is a positive correlation between the P-wave velocity and hydrate saturation. Also shown are the results of the load-bearing model (Best et al., 2013), whose values are generally higher than the measured ones. For well 1250F, at the hydrate saturation range of 0.1–0.15, the average deviation of the triple-porosity model predictions with respect to the logging data is 22.84 m/s and that of the load-bearing model is 48.93 m/s.

FIGURE 12
www.frontiersin.org

FIGURE 12. Measured P-wave velocity data (symbols) compared to the results of the triple-porosity and load-bearing models varying with hydrate saturation in Wells 1247B (A) and 1250F (B).

Nankai-trough data

We consider the sonic-log and VSP data obtained by Zhan and Matsushima. (2018) in the Nankai Trough, Japan. The frequency is 14 kHz, the strata rock porosity is approximately in the range of ϕ = 35%–43%, the grain coordination number is n = 8.5, and the seawater viscosity is η = 0.0018 kg/(m•s). Figures 13A, B compare the measured and theoretical P-wave velocities and dissipation factor as a function of hydrate saturation, for the double- and triple-porosity models. The theoretical porosity is ϕ = 35%, the clay radius is R23 = 0.2 cm, and the hydrate inclusion radius is R12 = 0.075 cm. The velocity gradually increases with hydrate saturation, and the variation of the measured P-wave attenuation is relatively large, possibly related to different hydrate morphologies not considered here (see Figure 1). The first model predicts a higher velocity when the hydrate saturation exceeds 20%. The triple-porosity model shows a better agreement.

FIGURE 13
www.frontiersin.org

FIGURE 13. Measured P-wave velocity (A) and dissipation factor (B) (symbols) compared to the results of the double- and triple-porosity models.

Conclusion

The mechanisms of wave propagation in hydrate-bearing sediments are analyzed by using a triple-porosity model. Specifically, we obtain the P-wave velocity and attenuation as a function of frequency, inclusion radius of the clay and hydrate phases, porosity, and hydrate volume ratio. The model considers three attenuation mechanisms, namely, two due to local fluid flow between the rock frame and clay and hydrate inclusions (mesoscopic loss) and the classical global Biot loss. Local flow effects dominate at low (seismic) frequencies. Well-log data from ODP204 site and offshore Japan are compared to the model predictions, which show a good agreement.

Data availability statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

Author contributions

JB: modeling, writing and verification. FG: modeling and writing. JC: writing and verification. DG: writing and verification.

Funding

The authors were grateful to the support of the Jiangsu Innovation and Entrepreneurship Plan, research funds from SINOPEC Key Laboratory of Geophysics, Jiangsu Province Science Fund for Distinguished Young Scholars (BK20200021), and National Natural Science Foundation of China (41974123).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Ba, J., Carcione, J. M., Cao, H., Du, Q. Z., Yuan, Z. Y., and Lu, M. H. (2012). Velocity dispersion and attenuation of P waves in partially-saturated rocks: Wave propagation equations in double-porosity medium. Chinese J. Geophys. 55, 219–231. doi:10.6038/j.issn.0001-5733.2012.01.021

CrossRef Full Text | Google Scholar

Ba, J., Carcione, J. M., and Nie, J. X. (2011). Biot-Rayleigh theory of wave propagation in double-porosity media. J. Geophys. Res. 116, B06202. doi:10.1029/2010jb008185

CrossRef Full Text | Google Scholar

Ba, J., Zhao, J., Carcione, J. M., and Huang, X. X. (2016). Compressional wave dispersion due to rock matrix stiffening by clay squirt flow. Geophys. Res. Lett. 43, 6186–6195. doi:10.1002/2016gl069312

CrossRef Full Text | Google Scholar

Best, A. I., Priest, J. A., Clayton, C. R. I., and Rees, E. V. L. (2013). The effect of methane hydrate morphology and water saturation on seismic wave attenuation in sand under shallow sub-seafloor conditions. Earth Planet. Sci. Lett. 368, 78–87. doi:10.1016/j.epsl.2013.02.033

CrossRef Full Text | Google Scholar

Biot, M. (1962). Generalized theory of acoustic propagation in porous dissipative media. J. Acoust. Soc. Am. 34, 1254–1264. doi:10.1121/1.1918315

CrossRef Full Text | Google Scholar

Carcione, J. M., and Gei, D. (2004). Gas-hydrate concentration estimated from P- and S-wave velocities at the Mallik 2L-38 research well, Mackenzie Delta, Canada. J. Appl. Geophy. 56, 73–78. doi:10.1016/j.jappgeo.2004.04.001

CrossRef Full Text | Google Scholar

Carcione, J. M., Gei, D., Rossi, G., and Madrussani, G. (2005). Estimation of gas-hydrate concentration and free-gas saturation at the Norwegian-Svalbard continental margin. Geophys. Prospect. 53, 803–810. doi:10.1111/j.1365-2478.2005.00502.x

CrossRef Full Text | Google Scholar

Carcione, J. M. (2022). Wave fields in real media. Theory and numerical simulation of wave propagation in anisotropic, anelastic, porous and electromagnetic media. 4th Edn. Amsterdam: Elsevier.

Google Scholar

Chand, S., and Minshull, T. A. (2004). The effect of hydrate content on seismic attenuation: A case study for mallik 2L-38 well data, mackenzie delta, Canada. Geophys. Res. Lett. 31, L14609. doi:10.1029/2004gl020292

CrossRef Full Text | Google Scholar

Dai, J., Xu, H., Snyder, F., and Dutta, N. (2004). Detection and estimation of gas hydrates using rock physics and seismic inversion: Examples from the northern deepwater Gulf of Mexico. Lead. Edge 23, 60–66. doi:10.1190/1.1645456

CrossRef Full Text | Google Scholar

Dvorkin, J., Prasad, M., Sakai, A., and Lavoie, D. (1999). Elasticity of marine sediments: Rock physics modeling. Geophys. Res. Lett. 26, 1781–1784. doi:10.1029/1999gl900332

CrossRef Full Text | Google Scholar

Ecker, C., Dvorkin, J., and Nur, A. M. (2000). Estimating the amount of gas hydrate and free gas from marine seismic data. Geophysics 65, 565–573. doi:10.1190/1.1444752

CrossRef Full Text | Google Scholar

Ecker, C., Dvorkin, J., and Nur, A. M. (1998). Sediments with gas hydrates: Internal structure from seismic AVO. Geophysics 63, 1659–1669. doi:10.1190/1.1444462

CrossRef Full Text | Google Scholar

Gei, D., and Carcione, J. M. (2003). Acoustic properties of sediments saturated with gas hydrate, free gas and water. Geophys. Prospect. 51, 141–158. doi:10.1046/j.1365-2478.2003.00359.x

CrossRef Full Text | Google Scholar

Gei, D., Carcione, J. M., and Picotti, S. (2022). “Seismic rock physics of gas-hydrate bearing sediments,” in World atlas of submarine gas hydrates in continental margins. Editor J. Mienert (Springer), 55–63.

CrossRef Full Text | Google Scholar

Guerin, G., and Goldberg, D. (2005). Modeling of acoustic wave dissipation in gas hydrate-bearing sediments. Geochem. Geophys. Geosyst. 6, Q07010. doi:10.1029/2005gc000918

CrossRef Full Text | Google Scholar

Helgerud, M. B., Dvorkin, J., Nur, A., Sakai, A., and Collett, T. (1999). Elastic-wave velocity in marine sediments with gas hydrates: Effective medium modeling. Geophys. Res. Lett. 26, 2021–2024. doi:10.1029/1999gl900421

CrossRef Full Text | Google Scholar

Hill, R. (1952). The elastic behaviour of a crystalline aggregate. Proc. Phys. Soc. A 65, 349–354. doi:10.1088/0370-1298/65/5/307

CrossRef Full Text | Google Scholar

Hu, G. W., Ye, Y. G., Zhang, J., Liu, C. L., Diao, S. B., and Wang, J. S. (2010). Acoustic properties of gas hydrate–bearing consolidated sediments and experimental testing of elastic velocity models. J. Geophys. Res. 115, B02102. doi:10.1029/2008JB006160

CrossRef Full Text | Google Scholar

Lee, M., Hutchinson, D., Collett, I., and Dillon, W. P. (1996). Seismic velocities for hydrate-bearing sediments using weighted equation. J. Geophys Res. 101, 20347–20358. doi:10.1029/96jb01886

CrossRef Full Text | Google Scholar

Lee, M. W., and Collett, T. S. (2009). Gas hydrate saturations estimated from fractured reservoir at Site NGHP-01-10, Krishna-Godavari Basin, India. J. Geophys. Res. 114, B07102. doi:10.1029/2008JB006237

CrossRef Full Text | Google Scholar

Leurer, K. C., and Brown, C. (2008). Acoustics of marine sediment under compaction: Binary grain-size model and viscoelastic extension of Biot’s theory. J. Acoust. Soc. Am. 123, 1941–1951. doi:10.1121/1.2871839

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, C., Feng, K., and Liu, X. (2015). Retracted article: Analysis of P-wave attenuation in hydrate-bearing sediments in the shenhu area, south China sea. Mar. Geophys. Res. 36, 357. doi:10.1007/s11001-014-9241-9

CrossRef Full Text | Google Scholar

Liu, J., Liu, J, P., Cheng, F., Wang, J., and Liu, X, X. (2017). Rock-physics models of hydrate-bearing sediments in permafrost, Qilian Mountains, China. Appl. Geophys. 4 (1), 31–39. doi:10.1007/s11770-017-0608-y

CrossRef Full Text | Google Scholar

Liu, L., Zhang, X., and Wang, X. (2021). Wave propagation characteristics in gas hydrate-bearing sediments and estimation of hydrate saturation. Energies 14 (4), 804. doi:10.3390/en14040804

CrossRef Full Text | Google Scholar

Marín-Moreno, H., Sahoo, S. K., and Best, A. I. (2017). Theoretical modeling insights into elastic wave attenuation mechanisms in marine sediments with pore-filling methane hydrate. J. Geophys. Res. Solid Earth 122, 1835–1847. doi:10.1002/2016jb013577

CrossRef Full Text | Google Scholar

Matsushima, J. (2006). Seismic wave attenuation in methane hydrate-bearing sediments: Vertical seismic profiling data from the Nankai Trough exploratory well, offshore tokai, central Japan. J. Geophys. Res. 111, B10101. doi:10.1029/2005jb004031

CrossRef Full Text | Google Scholar

Pan, H. J., Li, H. B., Grana, D., Zhang, Y., Liu, T., Geng, C., et al. (2019). Quantitative characterization of gas hydrate bearing sediment using elastic-electrical rock physics models. Mar. Petroleum Geol. 105, 273–283. doi:10.1016/j.marpetgeo.2019.04.034

CrossRef Full Text | Google Scholar

Priest, J. A., Rees, E. V. L., and Clayton, C. R. I. (2009). Influence of gas hydrate morphology on the seismic velocities of sands. J. Geophys. Res. 114, B11205. doi:10.1029/2009jb006284

CrossRef Full Text | Google Scholar

Sahoo, S. K., North, L. J., Marín-Moreno, H., Minshull, T. A., and Best, A. I. (2019). Laboratory observations of frequency-dependent ultrasonic P-wave velocity and attenuation during methane hydrate formation in Berea sandstone. Geophys. J. Int. 219, 713–723. doi:10.1093/gji/ggz311

CrossRef Full Text | Google Scholar

Schicks, J. M., Naumann, R., Erzinger, J., Hester, K. C., Koh, C. A., and Dendy Sloan, E. (2006). Phase transitions in mixed gas hydrates: Experimental observations versus calculated data. J. Phys. Chem. B 110 (23), 11468–11474. doi:10.1021/jp0612580

PubMed Abstract | CrossRef Full Text | Google Scholar

Sloan, D. (1990). Clathrate hydrates of natural gases. Boca Raton, Fla: CRC Press.

Google Scholar

Stoll, D., and Bryan, G. (2009). Wave attenuation in saturated sediments. J. Acoust. Soc. Am. 47, 1440–1447. doi:10.1121/1.1912054

CrossRef Full Text | Google Scholar

Waite, W. F., Santamarina, J. C., Cortes, D. D., Dugan, B., Espinoza, D. N., Germaine, J., et al. (2009). Physical properties of hydrate-bearing sediments. Rev. Geophys. 47, RG4003. doi:10.1029/2008rg000279

CrossRef Full Text | Google Scholar

Wang, W., Ba, J., Carcione, J. M., Liu, X., and Zhang, L. (2021). Wave properties of gas-hydrate bearing sediments based on poroelasticity. Front. Earth Sci. (Lausanne) 9, 640424. doi:10.3389/feart.2021.640424

CrossRef Full Text | Google Scholar

Wood, A. W. (1955). A textbook of sound. New York: McMillan Co.

Google Scholar

Wood, A. W., Holbrook, W., and Hoskins, H. (2000). In situ measurements of P-wave attenuation in the methane hydrate-and agas-bearing sediments of the Blake Ridge. Proc. Odp. Sci. Results 164, 265–272. doi:10.2973/odp.proc.sr.164.246.2000

CrossRef Full Text | Google Scholar

Wyllie, M. J., Gregory, A. R., and Gardner, G. H. F. (1958). An experimental investigation of factors affecting elastic wave velocities in porous media. Geophysics 23, 459–493. doi:10.1190/1.1438493

CrossRef Full Text | Google Scholar

Zhan, L., Liu, B., Zhang, Y., and Lu, H. (2022). Rock physics modeling of acoustic properties in gas hydrate-bearing sediment. J. Mar. Sci. Eng. 10, 1076. doi:10.3390/jmse10081076

CrossRef Full Text | Google Scholar

Zhan, L. S., and Matsushima, J. (2018). Frequency-dependentP-wave attenuation in hydrate-bearing sediments: A rock physics study at Nankai Trough, Japan. Geophys. J. Int. 214, 1961–1985. doi:10.1093/gji/ggy229

CrossRef Full Text | Google Scholar

Zhang, R. W., Li, H. Q., and Wen, P. F. (2016). The velocity dispersion and attenuation of marine hydrate-bearing sediments. Chin. J. Geophys. 59 (9), 3417–3427. doi:10.6038/jcg20160924

CrossRef Full Text | Google Scholar

Zhang, L., Ba, J., and Yin, W. (2017). Seismic wave propagation equations of conglomerate reservoirs: A triple-porosity structure model. Chin. J. Geophys. 60 (3), 1073–1087. doi:10.6038/cjg20170320

CrossRef Full Text | Google Scholar

Zhang, L., Ba, J., and Carcione, J. M (2021). Wave propagation in infinituple-porosity media. Geophysical Research: Solid Earth 126, e2020JB021266. doi:10.1029/2020JB021266

CrossRef Full Text | Google Scholar

Zhang, L., Ba, J., Carcione, J. M, and Wu, C. (2022). Seismic wave propagation in partially saturated rocks with a fractal distribution of fluid-patch size. Journal of Geophysical Research: Solid Earth 127, e2021JB023809. doi:10.1029/2021JB023809

CrossRef Full Text | Google Scholar

Appendix A: The explicit expressions of the elastic parameters are (Zhang et al., 2017)

A=1ϕKs23GbKsKfQ1+Q2+Q3Q1=ϕ1β1Ksβ1+γQ2=ϕ2Ks1+γQ3=ϕ3Ksβ2γ+1R1=ϕ1Kfβ1/γ+1R2=ϕ2Kf1/γ+1R3=ϕ3Kf1/β2γ+1(A1)

where

γ=KsKfϕ1β1+ϕ2+ϕ3β21ϕKbKs(A2)

and

β1=ϕ20ϕ101Kh1ϕ10Kb11Km1ϕ20Kb2(A3)
β2=ϕ30ϕ201Km1ϕ20Kb21Kc1ϕ30Kb3(A4)

where Kb1, Kb2 and Kb3 are the bulk moduli of hydrate, rock, and clay frames, respectively, and Kh, Km and Kc are the bulk moduli of hydrate, minerals and clay, respectively.

The tortuosities of the three phases are

χ1=121+1ϕ10χ2=121+1ϕ20χ3=121+1ϕ30(A5)

Then, the density parameters are

ρ11=χ1ϕ1ρfρ22=χ2ϕ2ρfρ33=χ3ϕ3ρfρ01=ϕ1ρfρ11ρ02=ϕ2ρfρ22ρ03=ϕ3ρfρ33ρ00=ν11ϕ10ρh+ν21ϕ20ρm+ν31ϕ30ρcρ01ρ02ρ03(A6)

where ρm, ρh and ρc are the densities of minerals, hydrate and clay, respectively. Moreover,

b1=ϕ1ϕ10ηκ1(A7)
b2=ϕ2ϕ20ηκ2(A8)
b3=ϕ3ϕ30ηκ3(A9)

Appendix B

A plane-wave analysis is performed by substituting a time harmonic kernel ejωtkx (where ω is the angular frequency, k is the wave number vector, and x is the spatial variable vector) into Eqs 9a9fa–f9a9f (Zhang et al., 2017). The resulting dispersion equation is

a11k2+b11a12k2+b12a13k2+b13a14k2+b14a21k2+b21a22k2+b22a23k2+b23a24k2+b24a31k2+b31a32k2+b32a33k2+b33a34k2+b34a41k2+b41a42k2+b42a43k2+b43a44k2+b44=0(B1)

where

a11=A+2Gb+Q1ϕ2Q2ϕ1M012+Q2ϕ3Q3ϕ2M023a12=Q1+Q1ϕ2Q2ϕ1M112+Q2ϕ3Q3ϕ2M123a13=Q2+Q1ϕ2Q2ϕ1M212+Q2ϕ3Q3ϕ2M223a14=Q3+Q1ϕ2Q2ϕ1M312+Q2ϕ3Q3ϕ2M323a21=Q1+ϕ2R1M012a22=R1+ϕ2R1M112a23=ϕ2R1M212a24=ϕ2R1M312a31=Q2R2ϕ1M012ϕ3M023a32=R2ϕ1M112ϕ3M123a33=R21ϕ1M212+ϕ3M223a34=R2ϕ1M312+ϕ3M323a41=Q3ϕ2R3M023a42=ϕ2R3M123a43=ϕ2R3M223a44=R31ϕ2M323b11=ρ00ω2+jωb1+b2+b3b12=ρ01ω2jωb1b13=ρ02ω2jωb2b14=ρ03ω2jωb3b21=ρ01ω2jωb1b22=ρ11ω2+jωb1b23=b24=0b31=ρ02ω2jωb2b33=ρ22ω2+jωb2b32=b34=0b41=ρ03ω2jωb3b44=ρ33ω2+jωb3b42=b43=0(B2)
S12=ϕ1ϕ22R122ωρfω1/5+ϕ10/ϕ20+jη/5κ1+η/κ2ϕ103ϕ22R1ϕ12R2S23=ϕ3ϕ22R232ωρfω1/5+ϕ30/ϕ20+jη/5κ3+η/κ2ϕ303ϕ32R2ϕ22R3M012=Q1ϕ2Q2ϕ1/S12+ϕ1ϕ3R2Q2ϕ3Q3ϕ2/S12S231+ϕ1ϕ3R22/S12S23M112=ϕ2R1/S121+ϕ1ϕ3R22/S12S23M212=ϕ1R2/S12+ϕ1ϕ32R22/S12S231+ϕ1ϕ3R22/S12S23M312=ϕ1ϕ2ϕ3R2R3/S12S231+ϕ1ϕ3R22/S12S23M023=M012ϕ1ϕ3R2+Q2ϕ3Q3ϕ2S23M123=M112ϕ1ϕ3R2/S23M223=M212ϕ1ϕ3R2+ϕ3R2S23M323=M312ϕ1ϕ3R2ϕ2R3/S23(B3)

where η is fluid viscosity. j=1.

VP=ωRek(B4)
Q1=Imk2Rek2(B5)

Keywords: triple-porosity theory, attenuation, rock physics model, hydrate, dispersion

Citation: Ba J, Guo F, Carcione JM and Gei D (2023) P-wave anelasticity in hydrate-bearing sediments based on a triple-porosity model. Front. Earth Sci. 10:1097550. doi: 10.3389/feart.2022.1097550

Received: 14 November 2022; Accepted: 01 December 2022;
Published: 04 January 2023.

Edited by:

Lidong Dai, Institute of geochemistry (CAS), China

Reviewed by:

Zhen Li, Chengdu University of Technology, China
Boye Fu, Beijing University of Technology, China

Copyright © 2023 Ba, Guo, Carcione and Gei. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jing Ba, jba@hhu.edu.cn

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.