THE RELATIONSHIP BETWEEN THE DEPTH OF THE SULFATE- METHANE TRANSITION AND GAS HYDRATE OCCURRENCE IN THE NORTHERN CASCADIA MARGIN (IODP EXP.

Similar documents
Article Volume 12, Number 7 12 July 2011 ISSN:

Analytical theory relating the depth of the sulfate methane transition to gas hydrate distribution and saturation

EFFECTS OF HETEROGENEOUS LITHOLOGY AND FOCUSED FLUID FLOW ON GAS HYDRATE DISTRIBUTION IN MARINE SEDIMENTS

STRESS AND GAS HYDRATE-FILLED FRACTURE DISTRIBUTION, KRISHNA-GODAVARI BASIN, INDIA

DATA REPOSITORY ITEM APPENDIX: MODEL SET-UP. The model, the Biogeochemical Reaction Network Simulator (BRNS, Regnier et al.

Proceedings of the 6th International Conference on Gas Hydrates (ICGH 2008), Vancouver, British Columbia, CANADA, July 6-10, 2008.

ABSTRACT. Modeling Fluid Flow Effects on Shallow Pore Water Chemistry and Methane. Hydrate Distribution in Heterogeneous Marine Sediment

Short migration of methane into a gas hydrate-bearing sand layer at Walker Ridge, Gulf of Mexico

A transfer function for the prediction of gas hydrate inventories in marine sediments

SEISMIC STRUCTURE, GAS-HYDRATE CONCENTRATIONS, AND SLUMPING ALONG THE IODP X311 TRANSECT ON THE N. CASCADIA MARGIN

Accumulation of gas hydrates in marine sediments

Earth and Planetary Science Letters

GAS HYDRATE QUANTIFICATION BY COMBINING PRESSURE CORING AND IN-SITU PORE WATER SAMPLING TOOLS

G 3. AN ELECTRONIC JOURNAL OF THE EARTH SCIENCES Published by AGU and the Geochemical Society

Carbon cycling within the sulfate-methane-transition-zone in marine sediments from the Ulleung Basin

Data report: stress orientations from borehole breakouts, IODP Expedition 308, Ursa area, Mississippi Fan, Gulf of Mexico 1

Using ammonium pore water profiles to assess stoichiometry of deep remineralization processes in methanogenic continental margin sediments

Data report: a downhole electrical resistivity study of northern Cascadia marine gas hydrate 1

Oil & Natural Gas Technology

Seismic interpretation of gas hydrate based on physical properties of sediments Summary Suitable gas hydrate occurrence environment Introduction

Seawater recharge along an eastern bounding fault in Middle Valley, northern Juan de Fuca Ridge

JOINT SEISMIC/ELECTRICAL EFFECTIVE MEDIUM MODELLING OF HYDRATE-BEARING MARINE SEDIMENTS AND AN APPLICATION TO THE VANCOUVER ISLAND MARGIN

210 Pb xs (mbq/g) ± 31-2 ± ± 12 3 ± ± 5 3 ± ± 11 4 ± 1

Tutorial on Methane Hydrate. Presented by Ad Hoc Group on Methane Hydrate Research March 24, 2004

Gas hydrate on the northern Cascadia margin: regional geophysics and structural framework 1

IODP EXPEDITION 311: CASCADIA MARGIN GAS HYDRATES SITE U1327 SUMMARY

Sulfate profiles and barium fronts in sediment on the Blake Ridge: Present and past methane fluxes through a large gas hydrate reservoir

Electrical and geomechanical Properties of Natural Gas Hydratebearing Sediments from Ulleung Basin, East Sea, Korea

INDIAN CONTINENTAL MARGIN GAS HYDRATE PROSPECTS: RESULTS OF THE INDIAN NATIONAL GAS HYDRATE PROGRAM (NGHP) EXPEDITION 01

DO DISSOCIATING GAS HYDRATES PLAY A ROLE IN TRIGGERING SUBMARINE SLOPE FAILURES? A CASE STUDY FROM THE NORTHERN CASCADIA MARGIN

17. DATA REPORT: CARBONATE, ORGANIC CARBON, AND OPAL CONCENTRATIONS SOUTHWEST AFRICA MARGIN 1

DYNAMICS OF SHALLOW MARINE GAS HYDRATE AND FREE GAS SYSTEMS

Estimation of methane flux offshore SW Taiwan and the influence of tectonics on gas hydrate accumulation

Global inventory of methane clathrate: sensitivity to changes in the deep ocean

Passing gas through the hydrate stability zone at southern Hydrate Ridge, offshore Oregon

Ship heave effects while drilling: observations from Legs 185 & 188

Numerical modelling: The governing equation used in this study is: (K T ) c T H 0,

Methane Migration and Its Influence on Sulfate Reduction in the Good Weather Ridge Region, South China Sea Continental Margin Sediments

Relations between the thermal properties and porosity of sediments in the eastern flank of the Juan de Fuca Ridge

24. SEDIMENT GRAIN-SIZE CONTROL ON GAS HYDRATE PRESENCE, SITES 994, 995, AND 997 1

Assessing methane release from the colossal Storegga submarine landslide

Annu. Rev. Earth Planet. Sci : Copyright 2000 by Annual Reviews. All rights reserved

Diffusive Evolution of Gaseous and Hydrate Horizons of Methane in Seabed

Methane dynamics in Santa Barbara Basin (USA) sediments as examined with a reaction-transport model

CONSTRAINING THE GLOBAL INVENTORY OF METHANE HYDRATE IN MARINE SEDIMENTS

Appendix 10: Non-Potential of Natural Gas Hydrate Occurrence in Queen Charlotte Basin8

Early diagenesis in marine sediments

Rates of methanogenesis and methanotrophy in deep-sea sediments

Gas hydrate on the continental margin. Hitoshi Tomaru

GOM 2 : Prospecting, Drilling and Sampling Coarse-Grained Hydrate Reservoirs in the Deepwater Gulf of Mexico

GSA DATA REPOSITORY

RISING ARCTIC OCEAN TEMPERATURES CAUSE GAS HYDRATE DESTABILIZATION AND OCEAN ACIDIFICATION

A Transformation from Acoustic and Density Properties to Reservoir Properties applied to Krishna Godavari Basin, India

Gas Hydrate BSR and Possible Fluid Migration in the Nankai Accretionary Prism off Muroto

SUPPLEMENTARY INFORMATION

Gas hydrate-related sedimentary pore pressure changes offshore Angola

Effect of porosity and permeability reduction on hydrate production in marine sediments

Downloaded 11/20/12 to Redistribution subject to SEG license or copyright; see Terms of Use at

11. CARBONATE CEMENTS AND FLUID CIRCULATION IN INTRAPLATE DEFORMATION 1

Chapter 1: Acoustic and elastic properties of calcareous sediments siliceous diagenetic front on the eastern U.S.

8. ELASTIC PROPERTY CORRECTIONS APPLIED TO LEG 154 SEDIMENT, CEARA RISE 1

Data report: pore water nitrate and silicate concentrations for Expedition 320/321 Pacific Equatorial Age Transect 1

SCOPE 35 Scales and Global Change (1988)

Methane Cycling in Coral Reef Frameworks

Geophysical Journal International

48. A SUMMARY OF INTERSTITIAL-WATER GEOCHEMISTRY OF LEG 133 1

SUPPLEMENTARY INFORMATION

Methane hydrate rock physics models for the Blake Outer Ridge

Hydrogen Sulphide-Rich Hydrates and Saline Fluids in the Continental Margin of South Australia

Methods of gas hydrate concentration estimation with field examples

Introduction. Theory. GEOHORIZONS December 2007/22. Summary

8. MODEL, STABLE ISOTOPE, AND RADIOTRACER CHARACTERIZATION OF ANAEROBIC METHANE OXIDATION IN GAS HYDRATE BEARING SEDIMENTS OF THE BLAKE RIDGE 1

Bay of Bengal Surface and Thermocline and the Arabian Sea

Characterization of in situ elastic properties of gas hydrate-bearing sediments on the Blake Ridge

27. DATA REPORT: NUCLEAR MAGNETIC RESONANCE LOGGING WHILE DRILLING, ODP LEG 204 1

DESIGN & IMPLEMENATION WORKSHOP. Conceptual Network Design For The Regional Cabled Observatory

Geologic and Oceanographic Setting

Journal of Asian Earth Sciences

*To whom correspondence should be addressed: FAX: (886) ;

Destabilization of Carbon Dioxide Hydrates in the Ocean. Resherle Verna Department of Earth Science, University of Southern California

21. FORMATION EVALUATION OF GAS HYDRATE BEARING MARINE SEDIMENTS ON THE BLAKE RIDGE WITH DOWNHOLE GEOCHEMICAL LOG MEASUREMENTS 1

Weissel, J., Peirce, J., Taylor, E., Alt, J., et al., 1991 Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 121

C. Global carbon cycle in high flux settings D. Fully parameterize global carbon cycle using wells of opportunity E. Preconditioning of

Fluid-Rock Interactions and the Sub-Seafloor Biosphere

Announcements. Manganese nodule distribution

Sediment Acoustics LONG-TERM GOAL

Bruce C. Heezen Graduate Research Fellowship

IODP drilling and core storage facilities

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B05104, doi: /2007jb005200, 2008

Capillary effects on hydrate stability in marine sediments

Along-strike variations in underthrust sediment dewatering on the Nicoya margin, Costa Rica related to the updip limit of seismicity

Hill, I.A., Taira, A., Firth, J.V., et al., 1993 Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 131

Flemings, P.B., Behrmann, J.H., John, C.M., and the Expedition 308 Scientists Proceedings of the Integrated Ocean Drilling Program, Volume 308

Diagenetic formation of greigite and pyrrhotite in gas hydrate marine sedimentary systems

9. SIGNIFICANCE OF ANAEROBIC METHANE OXIDATION IN METHANE-RICH SEDIMENTS OVERLYING THE BLAKE RIDGE GAS HYDRATES 1

Predicting the occurrence, distribution, and evolution

Scientific Report of AMBER subproject WP B.4. Identification and quantification of submarine groundwater. discharge

Marine Sediments Chapter Four Chapter Overview Marine Sediments Approaching the bottom (Alvin) Marine Sediments Classification of Marine Sediments

Techniques in Exploration and Formation Evaluation for Gas Hydrates

Essentials of Oceanography

Transcription:

Proceedings of the 7th International Conference on Gas Hydrates (ICGH 2011), Edinburgh, Scotland, United Kingdom, July 17-21, 2011. THE RELATIONSHIP BETWEEN THE DEPTH OF THE SULFATE- METHANE TRANSITION AND GAS HYDRATE OCCURRENCE IN THE NORTHERN CASCADIA MARGIN (IODP EXP. 311) Alberto Malinverno Lamont-Doherty Earth Observatory of Columbia University Palisades, NY 10964 U.S.A. ABSTRACT In a transect of four sites drilled by IODP Exp. 311 across the northern Cascadia margin, the depth of the near-seafloor sulfate-methane transition (SMT) is correlated to the depth to the top of the gas hydrate-bearing interval below. This relationship is expected if anaerobic oxidation of methane (AOM) is the dominant sulfate reduction mechanism, so that the depth of the SMT is controlled by the diffusive upward flux of methane. Sulfate, however, can also be consumed above the SMT by organoclastic sulfate reduction. This paper quantifies the relative importance of these two sulfate reduction mechanisms with estimates of the methane concentration gradient from pore water data and reaction-transport modeling. The independently estimated methane gradients are consistent at each site and show that at least 1/3 to 2/3 of the sulfate is consumed by AOM. The similar fraction of sulfate consumed by AOM in the transect sites explains the observed correlation between the depth of the SMT and the top of gas hydrate occurrence. Keywords: gas hydrates, sulfate-methane transition, anaerobic oxidation of methane NOMENCLATURE AOM Anaerobic oxidation of methane C org Organic carbon in sediment [dry weight %] c' m Concentration gradient of methane [mm/m] c' s Concentration gradient of sulfate [mm/m] D m Diffusion coefficient of methane [m 2 /s] D s Diffusion coefficient of sulfate [m 2 /s] F s Fraction of sulfate consumed by AOM GHOZ Gas hydrate occurrence zone GHSZ Gas hydrate stability zone mbsf Depth below the seafloor [m] SMT Sulfate-methane transition INTRODUCTION A four-site transect drilled across the northern Cascadia continental margin during Expedition 311 of the Integrated Ocean Drilling Program [1] provides a unique data set to advance our understanding of marine gas hydrates (Figure 1). The three seaward sites of the transect (U1326, U1325 and U1327) contain hydrates in a gas hydrate occurrence zone (GHOZ) whose top deepens moving landward [2]. There is no clear evidence of gas hydrate in the most landward site (U1329). The transect sites also display a shallow sulfate-methane transition (SMT), where sulfate and methane concentrations approach zero. As the SMT also deepens systematically moving landward, the depths of the SMT and of the top of the GHOZ are correlated (Figure 2). Explaining this correlation is the focus of this paper. The relationship in Figure 2 is expected if pore water sulfate is mostly reduced by anaerobic oxidation of methane (AOM). In this case, sulfate will be consumed quickly if the upward diffusive flux of methane is high, with higher methane fluxes corresponding to shallower SMT depths. A relatively high methane flux requires a high Corresponding author: Phone: +1 845 365 8577 Fax +1 845 365 8156 E-mail: alberto@ldeo.columbia.edu

Figure 1. Sites drilled on the northern Cascadia margin during IODP Exp. 311 (U1325-U1329) and ODP Leg 146 (889 and 890). The four-site transect studied in this paper is indicated by a dashed gray line. concentration gradient of methane, which in turn implies a relatively shallow occurrence of gas hydrate [3]. Sulfate reduction, however, can also be coupled with oxidation of particulate organic matter in organoclastic sulfate reduction [4]. If this process dominates sulfate consumption, the depth of the SMT will not be related to the top of the GHOZ. If both sulfate reduction processes are active, a relationship such as that in Figure 2 may still hold if the fraction of sulfate consumed by AOM is similar in all sites. Hence, quantifying the relative contribution of AOM and organoclastic sulfate reduction is key to understanding the relationship between the depth of the SMT and the occurrence of gas hydrates below. The relative contribution of AOM and organoclastic sulfate reduction can be determined by comparing the diffusive fluxes of sulfate and methane at the SMT. If the methane flux matches the sulfate flux, then all sulfate must be consumed by AOM [e.g., 5]; if the methane flux is less than the sulfate flux, some sulfate must be consumed by organoclastic sulfate reduction [e.g., 6]. Prior modeling work estimated that AOM consumes about half of the sulfate at Site U1325 [7]. This paper extends the previous analysis across the northern Cascadia transect. The goal is to test if the observed correlation between the depths of the SMT and of the top of the GHOZ can be explained Figure 2. The depths to the top of the GHOZ and to the SMT are correlated in the three northern Cascadia transect sites that have gas hydrates (black squares). A projected depth for the top of the GHOZ at Site U1329 can be obtained by extrapolating a linear relationship fitted to the other three sites to the SMT depth of Site U1329 (gray square). This projected depth is near the base of the GHSZ (dashed gray line), in agreement with the lack of gas hydrates at Site U1329. by AOM consuming a similar fraction of sulfate across the whole transect. METHOD The fraction of sulfate F s consumed by AOM at each site is determined from the ratio of the diffusive fluxes of sulfate and methane at the SMT [e.g., 5] as in F s = (D m c' m ) / (D s c' s ), (1) where D m and D s are diffusion coefficients, c' m and c' s are concentration gradients, and the subscripts m and s denote methane and sulfate, respectively. The diffusion coefficients in [8] give a ratio D m / D s = 1.568, and the sulfate gradients are estimated from the sulfate concentration at the seafloor (assumed to be 29 mm as in typical seawater) divided by the depth to the SMT at each site (Table 1).

Site Sedim. rate (m/ma) SMT depth (mbsf) Depth to top GHOZ (mbsf) Sulfate gradient at SMT (mm/m) Methane gradient at SMT from pore water data (mm/m) Methane gradient at SMT from model (mm/m) Sulfate fraction consumed by AOM from pore water data Sulfate fraction consumed by AOM from model U1326 366 2.5 47 11.60 3.13 2.75 0.42 0.37 U1325 302 4.5 73 6.44 2.17 1.89 0.53 0.46 U1327 187 8.5 111 3.41 1.43 1.32 0.66 0.60 U1329 98 9.5 3.05 0.88 0.65 0.45 0.33 Table 1. Key parameters at each site in the northern Cascadia transect. Sedimentation rates are averages in the GHSZ after [9], depths of the SMT are after [1], and depths to the top of the GHOZ are after [2]. Concentration gradients and sulfate fractions consumed by AOM are estimated in this study. Two methods are used to estimate the methane concentration gradients, based on pore water chemistry analyses and on reaction-transport modeling. Methane gradients from pore water chemistry A fundamental difficulty in using pore water analyses is that the low atmospheric solubility of methane causes loss of methane in recovered core samples. The methane concentration gradients from pore water chemistry listed in Table 1 were obtained from the first few samples immediately below the SMT that had concentrations below atmospheric solubility (~2 mm, [5]). Only 2-3 measured pore water concentrations could be typically used at each site, so that these methane gradients have large uncertainties and are likely underestimated. Methane gradients from reaction-transport modeling Methane concentration gradients at the SMT were estimated by applying the analytical reactiontransport model in [10] for a uniform sediment lithology. This model assumes steady-state conditions, constant porosity, and no upward water advection. The only source of methane is microbial in situ production. Modeling calculates a methane concentration profile whose gradient at the SMT can be used to estimate the fraction of sulfate F s consumed by AOM (Equation 1). This methane gradient estimate is likely a minimum because it does not account for upward pore water advection, which can cause a steeper increase in methane concentration with depth [2]. Key parameters used in the modeling at each site are listed in Table 1. At all sites, modeling used an average porosity of 50%, a tortuosity of 2.4 to compute the diffusion coefficients in bulk sediment, and a methane solubility profile computed from the local temperature gradient following [11]. The parameters that control in situ microbial methane production and the resulting methane concentration profile are the amount of metabolizable organic carbon (C org ) at the SMT and the reaction rate of methanogenesis. The amount of metabolizable C org can be estimated as the difference between the observed C org at the SMT and the amount that remains at depth, assumed to be the refractory fraction [12]. A conservative estimate for the metabolizable C org content in the northern Cascadia transect sites is a range 0.25% to 0.5% (dry weight fraction). Previous modeling studies in the Cascadia margin estimated a methanogenesis reaction rate on the order of 10 13 s 1 [4, 13]. The rest of this section describes the method used to estimate the methanogenesis reaction rate at each site. Sites U1325, U1326, and U1327 Gas hydrates have been detected at these sites, and the depth to the top of the GHOZ constrains the methanogenesis rate for a given metabolizable C org content [10]. In practice, a methanogenesis rate that is too low will mean that the depth where the modeled methane concentration reaches the solubility and gas hydrates start to form will be deeper than the observed top of the GHOZ. Conversely, if the rate is too high, the predicted

Figure 3. (A) Contour map of the depth where modeled methane concentration reaches solubility at Site U1327 for a given methanogenesis reaction rate and C org content at the SMT. The gray contour is a depth 5 m above the top of the GHOZ at this site. The two black dots mark the methanogenesis reaction rates that fit the observed top of the GHOZ for C org contents of 0.25% and 0.5%. (B) Methane concentration profile corresponding to the values of reaction rate and Corg content indicated by the dashed gray arrow in Figure 3A. The modeled methane concentration profile (solid black line) gives a gradient at the SMT indicated by the dotted black line. top of the GHOZ will be too shallow. The methanogenesis rates estimated at these sites are the rates that make the methane concentration reach solubility at a depth just above the observed top of the GHOZ for the assumed range of metabolizable C org content (0.25%-0.5%; Figure 3). Two values of the metabolizable C org content and of methanogenesis rates result in two values for the methane concentration gradient at the SMT; the differences in gradient, however, are small (at most 11% of the average). The average concentration gradients are listed in Table 1. Site U1329 There is no clear evidence of gas hydrates and no GHOZ at Site U1329, so the method used for the other three sites cannot be applied. Instead, methanogenesis rates at this site are predicted on the basis of their relationship to sedimentation rate. Methanogenesis rates have been proposed to be proportional to the square of the sedimentation rate [12], with higher sedimentation rates resulting in better preservation of organic matter. The methanogenesis rates estimated at the three transect sites with gas hydrates are indeed proportional to the square of the sedimentation rate (Figure 4). An extrapolation of the best-fit power law relationship to the sedimentation rate at Site U1329 gives predicted methanogenesis rates. Modeling results are consistent with the observed lack of gas hydrates at Site U1329, because the methane concentration calculated from these predicted rates does not reach saturation in the GHSZ of Site U1329. RESULTS AND DISCUSSION The methane concentration gradients estimated independently from pore water data and reactiontransport modeling are similar at each site (Table 1). Methane gradients from modeling are lower

Figure 4. The methanogenesis reaction rates estimated at the three northern Cascadia transect sites that have gas hydrates (filled circles) are approximately proportional to the square of the sedimentation rate. Best-fit power laws (dashed lines) have exponents of 2.17 and 2.2. Extrapolating the best-fit power laws to the sedimentation rate of Site U1329 give estimated methanogenesis reaction rates at this site (white circles). than those from pore water data, possibly because the modeling did not account for upward pore water advection. The fraction of sulfate consumed by AOM estimated in the transect sites varies between 1/3 and 2/3 (Table 1). At least part of this variation in sulfate fractions may reflect uncertainties in the estimated concentration gradients. The estimated sulfate gradients have uncertainties due to nonlinear concentration trends with depth and to uncertainties in the depth to the SMT. In most sites, methane concentration gradients from pore water measurements are computed from only 1-2 samples with methane concentrations below atmospheric. Hence, even a small error in sample depth or measured concentration can significantly affect the estimated gradient. Uncertainties of the methane gradients from modeling are also likely to be significant because of uncertainties in sedimentation rate and because pore water advection was neglected. These uncertainties can lead to changes in the estimates of the sulfate fraction consumed by AOM made at any one site that are as large as the total variation across the whole transect. For example, consider the fraction of sulfate consumed by AOM at Site U1325. An error of only 20 cm in the depth of the sample below the SMT that controls the methane gradient from pore water data gives a range of estimated sulfate fractions of 0.42 to 0.7. Given these uncertainties, the results in Table 1 are consistent with a fraction of sulfate consumed by AOM that is about the same in all the transect sites and is approximately equal to the fraction consumed by organoclastic sulfate reduction. A similar fraction of sulfate consumed by AOM at all sites explains the observed relationship between the depth of the SMT and the depth of the top of the GHOZ in the northern Cascadia transect (Figure 2). As the estimated methane gradients are likely to be lower bounds, the estimated fractions of sulfate consumed by AOM are minimum values and could be higher. The relationship between depth of the SMT and gas hydrate occurrence reported here is not universal. For example, no such relationship has been observed in sites drilled on the Bengal sea margin of India [14]. A possible explanation is that in this setting the fraction of sulfate consumed by AOM is more variable than in the northern Cascadia transect. CONCLUSIONS In a transect of sites drilled across the northern Cascadia margin, the depth of the near-seafloor sulfate-methane transition (2.5-9.5 mbsf) is correlated to the depth to the top of the gas hydrate-bearing interval below (47-111 mbsf). This relationship is expected if anaerobic oxidation of methane dominates sulfate reduction or consumes a similar fraction of sulfate at all sites. Methane concentration gradients estimated independently from pore water data and reactiontransport modeling show that at least 1/3 to 2/3 of the sulfate is consumed by anaerobic oxidation of methane in the transect sites. The similar fraction of sulfate consumed by AOM across the transect explains the observed correlation between the depths of the SMT and of gas hydrate occurrence. Modeling also shows that microbial methanogenesis reaction rates are proportional to the square of the sedimentation rate, as previously suggested [12].

Acknowledgments Thanks to John Pohlman for many discussions and helpful comments on the manuscript. This research used samples and data provided by the Integrated Ocean Drilling Program (IODP). The efforts of the JOIDES Resolution shipboard and drilling personnel and of the IODP Expedition 311 scientific party are gratefully acknowledged. REFERENCES [1] Riedel, M., Collett, T. S., Malone, M. J., and Expedition 311 Scientists (2006). Proceedings IODP, Exp. 311. Integrated Ocean Drilling Program Management International, Inc. [2] Malinverno, A., Kastner, M., Torres, M. E., and Wortmann, U. G. (2008). Gas hydrate occurrence from pore water chlorinity and downhole logs in a transect across the northern Cascadia margin (Integrated Ocean Drilling Program Expedition 311). J. Geophys. Res., 113:B08103, doi:10.1029/2008jb005702. [3] Borowski, W. S., Paull, C. K., and Ussler, III, W. (1996). Marine pore-water sulfate profiles indicate in situ methane flux from underlying gas hydrate. Geology, 24:655 658. [4] Claypool, G. E., Milkov, A. V., Lee, Y.-J., Torres, M. E., Borowski, W. S., and Tomaru, H. (2006). Microbial methane generation and gas transport in shallow sediments of an accretionary complex, southern Hydrate Ridge (ODP Leg 204), offshore Oregon, USA. In Tréhu, A. M., Bohrmann, G., Torres, M. E., and Colwell, F. S., editors, Proc. ODP, Sci. Results, 204, p. 1 52. Ocean Drilling Program, College Station, TX. [5] Niewöhner, C., Hensen, C., Kasten, S., Zabel, M., and Schulz, H. D. (1998). Deep sulfate reduction completely mediated by anaerobic methane oxidation in sediments of the upwelling area off Namibia. Geochim. Cosmochim. Acta, 62:455 464. [6] Treude, T., Niggemann, J., Kallmeyer, J., Wintersteller, P., Schubert, C. J., Boetius, A., and Jørgensen, B. B. (2005). Anaerobic oxidation of methane and sulfate reduction along the Chilean continental margin. Geochim. Cosmochim. Acta, 69:2767 2779, doi:10.1016/j.gca.2005.01.002. [7] Malinverno, A. and Pohlman, J. W. (2011). Modeling sulfate reduction in methane hydratebearing continental margin sediments: Does a sulfate-methane transition require anaerobic oxidation of methane? Geochem., Geophys., Geosyst., in review. [8] Schulz, H. D. (2006). Quantification of early diagenesis: Dissolved constituents in pore water and signals in the solid phase. In Schulz, H. D. and Zabel, M., editors, Marine Geochemistry, p. 73 124. Springer-Verlag, Berlin, 2nd edition. [9] Akiba, F., Inoue, Y., Saito-Kato, M., and Pohlman, J. W. (2009). Data report: Diatom and foraminiferal assemblages in Pleistocene turbidite sediments from the Cascadia margin (IODP Expedition 311), northeast Pacific. In Riedel, M., Collett, T. S., Malone, M. J., and Expedition 311 Scientists, editors, Proceedings IODP, Exp. 311. Integrated Ocean Drilling Program Management International, Inc. [10] Malinverno, A. (2010). Marine gas hydrates in thin sand layers that soak up biogenic methane. Earth Planet. Sci. Lett., 292:399 408, doi:10.1016/j.epsl.2010.02.008. [11] Davie, M. K., Zatsepina, O. Y., and Buffett, B. A. (2004). Methane solubility in marine hydrate environments. Mar. Geol., 203:177 184. [12] Tromp, T. K., Van Cappellen, P., and Key, R. M. (1995). A global model for the early diagenesis of organic carbon and organic phosphorus in marine sediments. Geochim. Cosmochim. Acta, 59:1259 1284. [13] Davie, M. K. and Buffett, B. A. (2003). Sources of methane for marine gas hydrate: Inferences from a comparison of observations and numerical models. Earth Planet. Sci. Lett., 206:51 63. [14] Kastner, M., Torres, M. E., Solomon, E. A., and Spivack, A. J. (2008). Marine pore fluid profiles of dissolved sulfate: Do they reflect in situ methane fluxes? Fire in the Ice, National Energy Technology Laboratory, U.S. Dept. of Energy, Summer 2008:6 8.