You are on page 1of 12

Journal of Asian Earth Sciences 149 (2017) 172–183

Contents lists available at ScienceDirect

Journal of Asian Earth Sciences


journal homepage: www.elsevier.com/locate/jseaes

Full length article

Biogeochemical cycles at the sulfate-methane transition zone (SMTZ) and MARK


geochemical characteristics of the pore fluids offshore southwestern Taiwan

Ching-Yi Hua,b, Tsanyao Frank Yanga,1, George S. Burrc, Pei-Chuan Chuanga,d, ,
Hsuan-Wen Chena, Monika Waliaa, Nai-Chen Chena, Yu-Chun Huanga, Saulwood Line,
Yunshuen Wangf, San-Hsiung Chungf, Chin-Da Huangb,g, Cheng-Hong Chena
a
Department of Geosciences, National Taiwan University, Taipei, Taiwan
b
Exploration and Production Business Division, CPC Corporation, Taiwan
c
Department of Oceanography, National Sun Yat-Sen University, Kaohsiung, Taiwan
d
MARUM – Center for Marine Environmental Sciences University of Bremen, Bremen, Germany
e
Institute of Oceanography, National Taiwan University, Taipei, Taiwan
f
Central Geological Survey, MOEA, Taipei, Taiwan
g
Department of Earth Sciences, National Taiwan Normal University, Taipei, Taiwan

A R T I C L E I N F O A B S T R A C T

Dedication: This paper is dedicated to the In this study, we used pore water dissolved inorganic carbon (DIC), SO42−, Ca2+ and Mg2+ gradients at the
memory of the late Professor Tsanyao Frank sulfate-methane transition zone (SMTZ) to estimate biogeochemical fluxes for cored sediments collected offshore
Yang who contributed greatly to the study of SW Taiwan. Net DIC flux changes (ΔDIC-Prod) were applied to determine the proportion of sulfate consumption by
gas hydrates offshore SW Taiwan. His organic matter oxidation (heterotrophic sulfate reduction) and anaerobic oxidation of methane (AOM), and to
enthusiasm and constant encouragement to his
determine reliable CH4 fluxes at the SMTZ. Our results show that SO42− profiles are mainly controlled by AOM
graduate students made this study possible, and
he is sorely missed.
rather than heterotrophic sulfate reduction. Refinement of CH4 flux estimates enhance our understanding of
methane abundance from deep carbon reservoirs to the SMTZ. Concentrations of chloride (Cl−), bromide (Br−)
Keywords:
and iodide (I−) dissolved in pore water were used to identify potential sources that control fluid compositions
Methane
and the behavior of dissolved ions. Constant Cl− concentrations throughout ∼30 m sediment suggest no in-
Halogen
SMTZ fluence of gas hydrates for the compositions within the core. Bromide (Br−) and Iodine (I−) concentrations
Gas hydrate increase with sediment depth. The I−/Br− ratio appears to reflect organic matter degradation. SO42− con-
Offshore SW Taiwan centrations decrease with sediment depth at a constant rate, and sediment depth profiles of Br− and I− con-
centrations suggests diffusion as the main transport mechanism. Therefore diffusive flux calculations are rea-
sonable. Coring sites with high CH4 fluxes are more common in the accretionary wedge, amongst thrust faults
and fractures, than in the passive continental margin offshore southwestern Taiwan. AOM reactions are a major
sink for CH4 passing upward through the SMTZ and prevent high methane fluxes in the water column and to the
atmosphere.

1. Introduction are constant in seawater and active AOM curbs high methane fluxes
from below the SMTZ (Borowski et al., 1996). HS− is produced as a
Earth's largest CH4 reservoir is in marine sediments, primarily in gas metabolic waste product of AOM. Upward diffusion and oxidation of
hydrates (Kvenvolden, 1993; Milkov, 2004; Buffet and Archer, 2004), HS− can supply large amounts of energy to chemosynthetic benthic
however relatively minor amounts of CH4 are released from marine communities living on the sea floor and within sediments (Suess et al.,
sediments to the ocean and atmosphere due to efficient anaerobic 1985; Treude et al., 2003; Dale et al., 2010). In addition, the high flux
oxidation of methane (AOM) through sulfate reduction (Barnes and of (bacterial/thermogenic) CH4 fuels high microbial activity (inferred
Goldberg, 1976; Boetius et al., 2000; Borowski, 2004; Regnier et al., by high reaction rates) at gas seeps. These communities are a thriving
2011). AOM occurs in the sulfate-methane transition zone (SMTZ) and component of the deep biosphere. Indeed, AOM rates in seep sediments
both CH4 and SO42− are essential to the reaction. SO42− concentrations are higher than those in non-seep sediments and bacterial biomass and


Corresponding author at: MARUM – Center for Marine Environmental Sciences University of Bremen, Bremen, Germany.
E-mail addresses: pchuang@marum.de, pcchuang2@gmail.com (P.-C. Chuang).
1
Deceased.

http://dx.doi.org/10.1016/j.jseaes.2017.07.002
Received 30 June 2016; Received in revised form 1 July 2017; Accepted 2 July 2017
Available online 04 July 2017
1367-9120/ © 2017 Elsevier Ltd. All rights reserved.
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

activity are stimulated in specific regions of the hydrate zone (Cragg (2008). At the upper slope area, thrust faults and fractures become
et al., 1996; Suess et al., 2001; Joye et al., 2004; Orcutt et al., 2011). potential pathways for deep fluid migration and the site of submarine
CH4 is a natural greenhouse gas whose fluxes should be accurately mud volcanoes. The lower slope can be further divided into rear and
known to assess its influence on the global carbon budget, energy sto- frontal segments (Liu et al., 1997; Lin et al., 2008; Lin et al., 2009). In
rage and climate change. However it is not easy to obtain precise values general, the upper slope domain has NE-SW trending gullies and sub-
for CH4 fluxes in the ocean. Applying Fick’s First Law to estimate CH4 marine canyons, while the lower slope domain has N-S to NW-SE
flux is a direct method; but it requires precise depth profiles of CH4 trending ridges and troughs formed by structural deformation and
concentrations without the influence of degassing during sample re- surface erosion (Yu and Song, 2000). The frontal segment is located
covery (Niewöhner et al., 1998; Borowski et al., 1996). Calculations between the deformation front and the rear segment. The anticline,
from SO42− gradients assuming one-to-one stoichiometric consumption blind thrust faults and Tainan Ridge are the dominant structures in the
by AOM have been applied to some gas hydrate systems but this ap- frontal segment. In the rear segment emergent thrust faults dominate
proach excludes the influence of heterotrophic sulfate reduction (HSR) the ridges, such as “Yungan Ridge” and “Good Weather Ridge” (Lin
(Niewöhner et al., 1998; Borowski et al., 1996; Dickens, 2001; Kastner et al., 2008; Lin et al., 2009).
et al., 2008). In this region, various geophysical and geochemical features were
HSR and AOM reactions can be expressed follows: reported, including: (1) numerous bottom simulating reflectors (BSRs)
(e.g., Chi et al., 1998; Schnurle et al., 1999; Chow et al., 2000; Liu et al.,
Heterotrophic sulfate reduction (HSR): 2006; Jiang et al., 2006; Schnürlea et al., 2011; Chen et al., 2014a,
2014b), (2) high CH4 concentrations in bottom water and shallow se-
2(CH2 O) + SO24− → 2HCO−3 + H2 S (R1) diments (Chuang et al., 2006; Lin et al., 2006; Yang et al., 2006; Chuang
et al., 2010; Chuang et al., 2013) and (3) a shallow SMTZ, 1–3 m below
Anaerobic oxidation of methane (AOM): the seafloor (Chuang et al., 2010; Lim et al., 2011). All these are sig-
natures for the potential existence of gas hydrates and huge methane
CH 4 + SO24− → HCO−3 + HS− + H2 O (R2) reserves at depth.

Flux calculations from DIC can be used to make a comprehensive 3. Materials and analytical methods
flux estimate by considering the net change of the DIC flux (ΔDIC-Prod)
and SO42− flux across the SMTZ (Snyder et al., 2007; Wehrmann et al., Eleven sediment cores (about 30 m long) were collected with a
2011; Komada et al., 2016). In the case of our geochemical data off- Calypso piston corer on the Marion Dufresne 178 cruise (2010/05/
shore SW Taiwan, we estimate CH4 flux up to the SMT and quantify the 27–2010/06/08). These coring sites were selected from various geolo-
amount of SO42− reduction utilizing organic matter or CH4 by means of gical regions and include: (1) low Fanliao basin (site 3290), (2) lower
SO42−, DIC, Ca2+ and Mg2+ gradients. The approach we use in this slope (site 3291), (3) Yungan Ridge (site 3274 and 3277), (4) Good
study provides a more precise value than obtained with CH4 and SO42− Weather Ridge (site 3293), (5) Palm Ridge (site 3261, 3287, and 3288),
gradients alone. (6) Formosa Ridge (site 3262), (7) Frontal Ridge (site 3266), and (8) a
In addition to SO42− fluxes, pore water composition is another relatively tectonic stable area (site 3264) (Fig. 1 and Table 1).
important piece of information at cold seep environments to reveal the After core retrieval, sediment sections (6 and 7 cm long) were
behavior of underlying gas hydrates, deep biosphere processes, and sub–sampled at 50 cm intervals. Samples for hydrocarbon gas analysis
fluid sources (e.g., Torres et al., 2002, 2004a, 2004b; Wallmann et al., were obtained from 6 cm-long wet sediment samples by using plastic
2006a; Hiruta et al., 2009; Kastner et al., 2008; Lu et al., 2008; Luo 5 mL syringes with the needle attachment end removed. The sediment
et al., 2016). In addition, CH4, Br− and I− are released from buried (15 mL) plugs from the syringes were immediately extruded into 30 mL
sedimentary organic matter decomposition. Because sedimentation rate glass serum bottles, filled with saturated sodium chloride solution and
is invariably faster than the decomposition rate of sedimentary organic sealed with blue butyl stoppers and aluminum crimp caps. Samples for
matter, biodegraded Br− and I− increase gradually with depth (e.g., pore water chemistry analysis were collected from 7 cm-long wet se-
Peru Margin of Martin et al. (1993), Blake Ridge of Egeberg and diment samples. Sediment pore fluids were extracted onboard by
Dickens (1999), Hydrate Ridge of Fehn et al. (2006), and Nankai squeezing and filtered through a 0.22-µm nylon membrane syringe
Trough of Muramatsu et al. (2007)). Cl− is a conservative ion in organic filter. Pore waters were then split into two to three 5 mL polypropylene
matter biodegradation, and comparisons of Cl− with Br− and I− are vials without a headspace for DIC concentration analysis and two bot-
used here to assist in understanding fluid sources in the study area. tles for pore water anion and cation concentration analysis. The pore
Through the examination of pore water solute distributions within water samples for cation analysis were added to 0.1 mL 8 N nitric acid
sediments from 11 long cores (∼30 m), we enhance our understanding to prevent redox and precipitation reactions. Pore water samples were
of the processes affecting biogeochemical fluxes at the SMTZ, the fac- preserved at 4 °C until laboratory analysis.
tors controlling sources and sinks of pore water SO42−, CH4 and DIC, Samples for hydrocarbon gas analysis were determined using a gas
and the role of gas hydrates in this dynamic system. chromatograph. Samples were introduced into a GC equipped with a
4 m-long Hayesep D column in line with a helium ionization detector.
2. Study area The temperature scheme for gas separation proceeded with the initial
injection at 50 °C, being held for 9 min, and ramped to 200 °C at a rate
The study area is offshore southwestern Taiwan (Fig. 1), located of 90 °C min−1. Analytical precision is typically better than 5%.
along an arc-continent collision boundary between the Luzon arc of the Detailed headspace methods for gas extractions and gas chromato-
Philippine Sea Plate and the Chinese continental margin (Lin et al., graphic analysis for C1–C6 hydrocarbons have been documented pre-
2008; Lin et al., 2009). The collision has produced a number of pro- viously (Chuang et al., 2006, 2010; Sun et al., 2010). DIC pore water
minent north-south oriented geological structures, including the Manila concentrations were determined with an OI Analytical total organic
Trench, North Luzon Arc, and North Luzon Trough (Teng, 1990; Huang carbon (TOC) analyzer combined with a Picarro G1101–i cavity ring
et al., 1997; Liu et al., 1998; Liu et al., 2004). A deformation front down spectrometer (CRDS) isotopic analyzer. A total of 10–15 mL of
marks the boundary between accretionary wedge, with numerous water sample was treated with 5% H3PO4 in a glass vial at 25 °C on line.
ridges in the east, and passive continental margin in the west (Liu et al., The CO2 produced was stripped with N2 and introduced into the de-
1997; Fig. 1). The accretionary wedge consists of upper slope, lower tectors. The analytical errors are better than 5% for DIC concentration.
slope, and ridges, as described by Reed et al. (1992) and Lin et al. Among 11 cores, DIC data were obtained from eight cores chosen to

173
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 1. Bathymetric map offshore southwestern Taiwan, the boundary


of the upper and lower slope defined by Reed et al. (1992), and the
deformation front defined by Lin et al. (2008). Approximately 30 m
long sediment cores (triangles) were collected from different geologic
settings including the low Fanliao basin (site 3290), lower slope (site
3291), Yungan Ridge (site 3274 and 3277), Good Weather Ridge (site
3293), Palm Ridge (site 3261, 3287, and 3288), Formosa Ridge (site
3262), Frontal Ridge (site 3266), and a relatively tectonically stable
area (site 3264).

represent various geological structural settings. In this study, values in seawater of Boudreau (1997) were used
Pore water anion concentrations (Cl−, Br−, I− and SO42−) and (0.018819, 0.017660, 0.013185, and 0.012238 for DIC, SO42−, Ca2+,
cation concentrations (Ca2+ and Mg2+) were measured with ion and Mg2+, respectively). We used n = 3 for clay-silt sediments after
chromatography (882 Compact IC). The anion and cation aliquots were Niewöhner et al. (1998) and Chuang et al. (2010) and average porosity
diluted 200-fold and 100-fold, separately, with Milli-Q water to bring data (Chen, 2011).
these concentrations within the appropriate analytical range for the ion The net change of DIC flux across the SMTZ (ΔDIC-Prod) is calculated
chromatograph. The analytical errors of Cl−, Br−, I−, SO42−, Ca2+ and by Eq. (3):
Mg2+ were 4.1%, 3.3%, 3.3%, 4.0%, 2.0% and 1.0% respectively.
ΔDIC − Prod = FDIC − Shallow−FDIC − Deep−Fcarb (3)

4. Estimating biogeochemical fluxes where FDIC-Shallow and FDIC-Deep represent upward flux of DIC above and
below SMTZ, respectively. Fcarb is the difference in the sum of Ca2+,
CH4 fluxes at each coring site were calculated from biogeochemical and Mg2+ downward flux (FCa and FMg) below and above the SMTZ
DIC fluxes, SO42−, Ca2+, and Mg2+ as described by Wehrmann et al. (Wehrmann et al., 2011).
(2011). These biogeochemical fluxes were estimated according to Fick's Fcarb = FCa + Mg − Shallow−FCa + Mg − Deep (4)
First Law, assuming steady-state conditions, using Eq. (1):
where FCa+Mg-shallow and FCa+Mg-deep represent the sum of the Ca flux 2+
F= −ϕ·DS ·dc/dx (1)
and Mg2+ flux above and below the SMTZ.
where F is the diffusive flux (mmol m−2 year−1), ϕ is the porosity For a calculation method comparison, CH4 flux was also calculated
(using mean porosity over the depths with steep CH4 gradients; (Chen, from CH4 concentration data around the SMTZ using Fick’s First Law
2011)), c is the concentration of the dissolved element (mM), x is the (Eq. (1)). A tracer diffusion coefficient of 0.027436 (m2 year−1) was
depth (m) and DS is the sediment diffusion coefficient (m2 year−1). DS used, after Boudreau (1997).
can be derived from Eq. (2):
5. Results
DS = D0 /[1 + n(1−ϕ)] (2)

where D0 is the tracer diffusion coefficient of various ions (m2 year−1). The concentration of major dissolved ions (SO42−, Ca2+, Mg2+,

Table 1
Locations of piston coring (Calypso piston corer) during the Marion Dufresne 178 cruise.

a
Core ID Longitude (E) Latitude (N) Water depth (m) Length of barrel (m) Length of core (m) Geological Structure

MD10-3261 119°38.020′ 22°25.830′ 1520 23 19.52 Palm Ridge


MD10-3262 119°17.442′ 22°06.402′ 1200 35 29.2 Formosa Ridge
MD10-3264 119°19.630′ 21°32.490′ 2820 23 21.84 tectonically stable
MD10-3266 119°47.640′ 21°43.970′ 2890 23 22.66 Frontal Ridge
MD10-3274 119°53.570′ 22°14.450′ 1327 25 24.7 Yungan Ridge
MD10-3277 119°54.220′ 22°14.460′ 1395 25 25.68 Yungan Ridge
MD10-3287 119°41.760′ 22°29.230′ 1050 25 24.65 Palm Ridge
MD10-3288 119°45.040′ 22°26.600′ 1300 35 31.26 Palm Ridge
MD10-3290 120°20.140′ 21°53.000′ 1272 25 15.65 Lower Fangliao Basin
MD10-3291 120°13.910′ 21°41.490′ 2070 35 38.5 lower slope
MD10-3293 120°01.540′ 22°16.630′ 1004 25 22.72 Good Weather Ridge

a
Core ID are not continuous.

174
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 2. Coring sites with DIC data at accretionary wedge, depth profiles of concentrations of major dissolved ions (SO42−, Ca2+, Mg2+, Cl−, Br−, I−), DIC and dissolved methane. SMTZ
depths are represented by the horizontal gray bars. Seawater value is plotted by the star at 0 m below the seafloor. Numbers adjacent represent flux values (mmol m−2 year−1) of SO42−,
Ca2+, Mg2+, and DIC below and above the SMTZ. These fluxes are calculated from Fick’s First Law and concentration gradient (mM m−1).

175
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 2. (continued)

176
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 3. Coring sites with DIC data at passive continental margin, depth profiles of concentrations of major dissolved ions (SO42−, Ca2+, Mg2+, Cl−, Br−, I−), DIC and dissolved CH4.
SMTZ depths are represented by the horizontal gray bars. Seawater value is plotted as a star at 0 m below seafloor. Numbers adjacent allows represent flux values (mmol m−2 year−1) of
SO42−, Ca2+, Mg2+, and DIC below and above SMTZ. These fluxes are calculated from Fick’s First Law and concentration gradient (mM m−1).

Cl−, Br−, I−), CH4 and DIC concentrations are plotted against sediment (from 10 to 3 and 55 to 30 mM, respectively). Below the SMTZ, they
depth in Figs. 2 and 3. Seawater values are plotted at 0 m for reference maintain constant values. Therefore, the sum of Ca2+ and Mg2+ fluxes
(Dickson and Goyet, 1994). The depth profiles of Cl−, I−, and Br− below the SMTZ (FCa+Mg-deep) can be considered as 0 (Fcarb = FCa+Mg-
concentrations exhibit similar trends. Cl− concentration ranges be- shallow) in this study area. These profiles also indicate that Ca
2+
and
2+
tween 500 mM and 550 mM with no apparent deviations throughout. Mg concentrations are controlled by the same mechanisms affecting
These behaviors are commonly observed in gas hydrate areas (Martin sulfate, methane and DIC. The plot of DIC (cation-adjusted) added
et al., 1993; Egeberg and Dickens, 1999; Fehn et al., 2006; Muramatsu versus SO42− removed in Fig. 4 support AOM (R2) and HSR (R1) as the
et al., 2007). Most of the SO42− concentration depth profiles decrease possible mechanisms, since most of data fall within a 1:1–1:2 ratio in-
from the average seawater at the sediment surface (30 mM) to 0 mM terval. Since DIC is produced through AOM (R2) and HSR (R1), a si-
(below detection limit) at depth. CH4 is the dominant hydrocarbon gas multaneous decrease in calcium and magnesium concentrations sug-
and CH4 concentrations range from 0.017 mM to 1.67 mM of most sites gests either (1) Mg-rich calcite or (2) simultaneous dolomite and calcite
reach up to 0.7 mM except MD10-3261, C2+ gases were under detec- formation within the SMTZ (Kelts and McKenzie, 1982; Meister et al.,
tion limits for most of the cores. The depth of the SMTZ positioned by 2007; Ussler and Paull, 2008).
depleting concentrations of SO42− and CH4 is generally above 1 m. DIC
concentrations range from 0 to 25 mM. The concentration increases
from the sediment surface to the SMTZ, and then becomes depleted or
constant with sediment depth below the SMTZ.
Seawater Ca2+ and Mg2+ concentrations decrease toward the SMTZ

177
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

et al., 2006; Kastner et al., 2008; Chatterjee et al., 2011). However, both
approaches are not able to address the fractions of SO42− consumption
contributed from HSR and AOM, quantitatively.
Hence, we quantify SO42− fluxes via HSR (FSO4-HSR) and AOM (FSO4-
AOM) by solving the following two equations.

FSO4 − AOM + FSO4 − HSR = total SO24−flux (5)

FSO4 − AOM + 2FSO4 − HSR = total change of the DIC flux (ΔDIC − Prod) (6)

Based on the different stoichiometry of DIC production and SO42−


consumption from HSR and AOM, we assume that the fluxes of SO42−
reduction and DIC production via the AOM reaction both are FSO4-AOM
and the fluxes of SO42− reduction and DIC production via HSR are FSO4-
HSR and 2FSO4-HSR. Hence the total SO4
2−
flux equals FSO4-AOM+FSO4-
HSR , and the total change of the DIC fluxes (ΔDIC-Prod) is equal to FSO4-
AOM+2FSO4-HSR. Two unknown parameters (FSO4-AOM and FSO4-HSR) in
the two equations (Eqs. (5) and (6)) can be solved by the estimated total
sulfate flux and ΔDIC-Prod. Hence, FSO4-AOM can be used to represent CH4
flux at the SMTZ. At coring sites where ΔDIC-Prod exceeds the SO42− flux
toward the SMTZ, 60 to 85% of the SO42− flux is due to AOM and true
CH4 fluxes at these sites are 2.6 to 34.7 mmol m−2 year−1 (Table 2).
This approach was applied to seven cores with DIC data to infer CH4
flux and the proportion of SO42− utilized by AOM and HSR (Fig. 5 and
Fig. 4. DIC (cation-adjusted) added versus SO42− removed (relative to seawater con- Table 2). The results indicate that the major sink for SO42− is AOM
centrations). Dashed lines indicate 1:1 and 1:2 ratios. The 1:1 ratio is consistent with rather than HSR. SO42− depletion is almost entirely controlled by AOM
SO42− reduction coupled with AOM. The 1:2 ratio is consistent with HSR. at the MD10-3262, MD10-3291 and MD10-3293 coring sites (Fig. 5),
and partly controlled by HSR, in addition to AOM, at MD10-3274,
6. Discussion MD10-3287, MD10-3288 and MD10-3290 (Fig. 5). Although some
coring sites (eg. MD10-3274, MD10-3287, MD10-3288 and MD10-
6.1. Biogeochemical fluxes 3290) have a linear sulfate profile, the contribution of HSR is not
negligible for DIC production (Table 2). This indicates that using SO42−
Results of estimated biogeochemical fluxes (SO42−, CH4 and DIC fluxes as a proxy for estimating CH4 fluxes according to the linear
(ΔDIC-Prod)) at the SMTZ are shown in Table 2. At the MD10-3262, SO42− gradient may overestimate CH4 flux. In this study, solving dif-
MD10-3291 and MD10-3293 coring sites (Formosa Ridge, and Good ferent rates of DIC produced in two SO42− reduction reactions based on
Weather Ridge, respectively), the ΔDIC-Prod is almost equal to the the calculated total changes of the DIC flux (ΔDIC-Prod) produces reliable
downward flux of sulfate from the seafloor (8.0, 10.1 and SO42− fluxes to determine CH4 fluxes. The value of the CH4 flux cal-
31.7 mmol m−2 year−1, respectively). This implies that SO42− here culated by Eqs. (5) and (6) is much higher than calculated by CH4
was reduced primarily through the AOM reaction, since SO42− reduc- concentration gradients obtained around the SMTZ (Table 3).
tion coupled to the AOM reaction produces DIC and consumes CH4, The derived CH4 fluxes offshore SW Taiwan reveal higher values in
both following a 1:1 stoichiometry. Therefore, CH4 flux is represented the accretionary wedge than in the passive continental margin. These
by the SO42− flux value. calculated CH4 fluxes are also high when compared with other gas
At the MD10-3274, MD10-3287, MD10-3288, and MD10-3290 hydrate areas and upwelling areas, such as Blake Ridge
coring sites (Yungan Ridge, Palm Ridge, Palm Ridge and Lower (18 mmol m−2 year−1) and Hydrate Ridge (7.2–7.6 mmol m−2 year−1)
Fangliao Basin respectively), ΔDIC-Prod is higher than the downward (Borowski et al., 1996; Dickens, 2001). Such high CH4 flux values at
SO42− flux from 8.9 to 75.2 mmol m−2 year−1 (Table 2). This indicates active margins indicate a broad distribution of methane-enriched
that HSR (R1) and AOM (R2) are both responsible for SO42− con- sources in the deep sediments, presumably following fold-and-thrust
sumption, since HSR produces more DIC than AOM with its 1:2 stoi- structures as a pathway for methane to migrate upward (Chuang et al.,
chiometry. 2013).
The approach of Wehrmann et al. (2011) that we applied in this
study is similar to plotting alkalinity or cation-adjusted DIC added, 6.2. The role of authigenic carbonate precipitation
versus sulfate removed (relative to seawater concentrations), which can
help identify which process dominates SO42− reductions (Claypool Due to alkalinity increases by DIC production, high CH4 flux areas

Table 2
Calculated diffusive fluxes of different ions (SO42−, Ca2+, Mg2+, and DIC) at the SMTZ assuming Fick’s fist Law (mmol m−2 year−1).

Core ID CH4 Flux SO42− Flux Ca2+ Flux Mg2+ Flux DIC Flux (DIC DIC Flux (DIC ΔDIC-Prod SO42− Flux SO42− AOM: HSR
(inferred) (sulfate (calcium (magnesium gradient above gradient below (AOM) Flux (HSR)
gradient) gradient) gradient) SMTZ) SMTZ)

MD10-3262 8.7 8.7 1.7 2.7 −2.2 1.4 −8.0 8.7 0.0 100:0
MD10-3274 11.3 13.2 2.7 2.9 −9.5 0.0 −15.1 11.3 1.9 86:14
MD10-3287 9.5 13.7 2.2 4.2 −6.7 4.8 −17.9 9.5 4.2 69:31
MD10-3288 34.7 54.9 12.8 24.5 −28.0 9.9 −75.2 34.7 20.2 63:37
MD10-3290 2.6 5.7 1.6 2.0 −4.0 1.3 −8.9 2.6 3.1 59:41
MD10-3291 10.4 10.4 2.4 3.2 −4.5 0.0 −10.1 10.4 0.0 100:0
MD10-3293 27.6 29.6 5.1 6.1 −14.8 5.7 −31.7 27.6 2.0 93:7

178
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 5. Distributions of the methane fluxes and ratios between AOM


reaction and HSR derived from Eqs. (4) and (5).

Table 3 multiplying Ca2+ and CO32− activities (Eq. (11)). If the value of
Comparison of CH4 fluxes through the SMTZ derived from methane concentration gra- IAPCalcite in pore water is higher than the value of Ksp-Calcite which is
dients and from DIC and sulfate concentration gradients (mmol m−2 year−1).
10−8.48 at 25 °C and standard atmospheric pressure, Ca2+ and CO32−
Core ID CH4 Flux (estimated from CH4 Flux (estimated from DIC and
would interact to precipitate CaCO3 (Atkins and De Paula, 2006). Since
methane gradient) sulfate gradients) dolomite rimmed by ankerite and Mg-rich calcite have been recognized
offshore SW Taiwan (Jiang, 2011) and solid solutions of magnesite and
MD10-3262 3.4 8.7 siderite have been found in cores MD10-3291 and MD10-3290 (Jiang,
MD10-3264 2.1 –
MD10-3266 2.7 –
2015), we calculated the IAP of dolomite and magnesite in pore water
MD10-3274 3.7 11.3 to examine the possibility of Mg2+ consumption following authigenic
MD10-3287 2.4 9.5 carbonate precipitation. The net reactions and Ksp of dolomite and
MD10-3288 2.2 34.7 magnesite are shown as Eqs. (12) and (13), while Eqs. (14) and (15)
MD10-3290 0.9 2.6
were used to calculate IAPDolomite and IAPMagnesite. Assuming equili-
MD10-3291 3.5 10.4
MD10-3293 2.8 27.6 brium between the solid-phase calcite and minimum pore water, the
CO32− concentration that is necessary for carbonate precipitation can
be calculated from the pore water Ca2+ activity with Eq. (16). Ca2+,
with active AOM are characterized by CH4-derived authigenic carbo- Mg2+ and CO32− activities are represented as [aCa2+], [aMg2+] and
nates (Greinert et al., 2001; Mazzini et al., 2006). Because precipitation [a CO32−] in the equations.
of such carbonates with Ca2+ and Mg2+ also consumes DIC, Ca2+ and
Ca2 + + CO32 − → CaCO3 K sp − Calcite = 10−8.48 (10)
Mg2+ fluxes were taken into account for the CH4 flux calculation from
ΔDIC-Prod. In this paragraph, we calculate the ion activity product (IAP)
IAPCalcite = [a Ca2 +][a CO2 −] (11)
(Moore et al., 2004; Wehrmann et al., 2011) in order to see if the au- 3

thigenic carbonate precipitation is reasonable at the SMTZ in the study


Ca2 + + Mg2 + + 2CO32 − → CaMgCO3 K sp − Dolomite = 10−16.52 (12)
area. The solubility product constant (Ksp) of a solid-phase is tem-
perature- and pressure-dependent. If IAP < Ksp, solution is under-
Mg2 + + CO32 − → MgCO3 K sp − Magnesite = 10−7.5 (13)
saturated with the mineral, the mineral dissociates; and if IAP > Ksp,
solution is oversaturated with the mineral it begins to precipitate. IAPDolomite = [a Ca2 +][a Mg2 +][a CO2 −]2
3 (14)
Concentrations and coefficients of ions in a reaction equation are re-
quired to calculate IAP. We neglect the effects of temperature and IAPMagnesite = [a Mg2 +][a CO2 −]
3 (15)
pressure on the equilibrium constant of calcite in the IAP in the cal-
culation, however this will be much less uncertain than the assumption K sp − Calcite
of equilibrium with calcite (Moore et al., 2004). The activity [a] of [a CO2 −] =
3
[a Ca2+] (16)
substances are calculated by concentrations and activity coefficients γa
according to the Eq. (7)–(9): The log IAP values of dolomite and magnesite calculated by CO32−,
Ca 2+
and Mg2+ activities in all cores are ∼−16.23 to −15.51 and
[a Ca2 +] = [Ca2 +] ∗γ Ca2 + (7)
∼−7.75 to −7.04, respectively (Fig. 6). Log IAPDolomite at all sites are
[a Mg2 +] = [Mg2 +] ∗γ Mg2 + higher than log Ksp-Dolomite which is −16.5 at 25 °C and standard at-
(8)
mospheric pressure (Moore et al., 2004). In these sulfate-methane rich
[a CO2 −] = [CO32 −] ∗γ CO2 − sediments, high Log IAPDolomite values suggest the possibility of dolo-
3 3 (9)
mite precipitation coupled to Ca2+ and Mg2+ depletions in pore water
2+ 2+ 2−
The activity coefficients of Ca , Mg and CO3 are 0.288, 0.248 at all coring sites which might be correlated with microbially-mediated
and 0.207 respectively (Schulz, 2006). Taking the calcite precipitation dolomite formation by sulfate-reducing bacteria (Vasconcelos et al.,
reaction in Eq. (10) as an example, IAPCalcite is calculated by 1995; Warthmann et al., 2000). In contrast, a log IAPMagnesite value that

179
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

magnesite in sediments from offshore SW Taiwan.


Based on the calculated results above, the slope changes across the
SMTZ shown on the depth profiles of calcium and magnesium in most
cores can be explained by consumption of both ions for carbonate
precipitation at the SMTZ.

6.3. Indicator of fluid sources

As relatively conservative ions, Cl−, Br− and I− concentrations not


only reflect the biodegradation of sedimentary organic matter but also
serve as proxies that can prove the existence of gas hydrates or other
fluids (Martin et al., 1993; Egeberg and Dickens, 1999; Muramatsu
et al., 2007; Fehn et al., 2007). We compare our results below with sites
from the published literature where fluid sources or gas hydrates are
believed to exist (Muramatsu et al., 2007; Fehn et al., 2007).
Cl− concentrations are close to seawater values and have no ob-
vious changes with depth, suggesting no influence of gas hydrates
within the core, and Br− and I− concentrations increase gradually with
depth (Figs. 2, 3 and 7). Br− and I− are released from sedimentary
organic matter into pore water though diagenesis (e.g. Wallmann et al.,
2006 b; Lu et al., 2008), and their diffusion coefficients are similar (Li
and Gregory, 1974). If organic matter is the main source of Br− and I−,
their release will adopt a constant ratio. The observed initial I−/Br−
values (i.e. ratios with low I− and Br− concentrations) in pore water
are similar to calculated I/Br ratios for marine organic matter, and the
depth dependence of I−/Br− ratios correspond to the degree of marine
organic matter degradation (Martin et al., 1993). However, the dis-
tribution and slope of Br− and I− concentration depth profiles vary in
some intervals and reflect other fluid flow, as well as organic matter
degradation. I−/Br− distributions obtained from hydrate fields of the
Nankai Trough area in Japan (Muramatsu et al., 2007) can be divided
into two parts, controlled either by biodegradation, or mixing with
other fluids (Fig. 7B and C). Relative concentrations of Br− and I− over
a gas hydrate on the Peru Margin (Fehn et al., 2007) suggest insignif-
icant fluid flow appearance as they follow a nearly linear trend with
depth (Fig. 7D). We argue that if Br− and I− concentrations are con-
trolled by biodegradation, the relationships between Br− and I− will
exhibit a linear slope, from which an initial ratio can be inferred to
determine the original I/Br ratio of the source marine organic matter. If
the distribution and slope for Br− and I− concentrations vary with
depth, fluid mixing is possible.
Br− vs. I− concentrations of 11 coring sites collected from different
geological regions in this study display linear correlations (i.e. domi-
nated by diffusion), and similar slopes, from which initial I−/Br− ratios
can be determined (Fig. 7A). The average value of the initial I−/Br−
ratio is 0.06, and this value estimates the I/Br ratio in marine organic
Fig. 6. Depth profiles of log IAP for (A) dolomite and (B) magnesite calculated using matter offshore SW Taiwan. Diffusion dominates sulfate profiles as
CO32−, Ca2+ and Mg2+ concentrations of pore water. Dashed lines represent log Ksp- well, which suggests that ion concentrations are mainly controlled by
Dolomite and log Ksp-Magnesite (−16.5 and −7.5 at 25 °C and standard atmospheric pres- similar biodegradation with insignificant fluid flow. This indicates that
sure).
our calculation of diffusive fluxes at the SMTZ are reasonable.

is higher than log Ksp-Magnesite (−7.5 at 25 °C and standard atmo- 7. Conclusions


spheric pressure (Siegel, 1975)) appears at partial sediment depth in the
study area. The appearance is restricted at sediment depths in excess of This study documents the concentration of major dissolved ions
∼10 m for cores MD10-3266, MD10-3274, MD10-3277, MD10-3287, (SO42−, Ca2+, Mg2+, Cl−, Br−, I−) and DIC in surface sediment pro-
MD10-3288, MD10-3290, MD10-3291 and MD10-3293. In cores MD10- files collected from the SW margin of Taiwan. We estimated biogeo-
3261 and 3262, the log IAPMagnesie value is lower than −7.5 chemical fluxes (SO42−, DIC, Ca2+ and Mg2+) at the SMTZ to quanti-
throughout, indicating that magnesite formation is improbable. The tatively determine the consumption of SO42− from HSR of particulate
occurrence of magnesite, high Mg-calcite and dolomite in other en- organic matter degradation and AOM, and obtain reliable CH4 fluxes.
vironments has been reported previously (e.g., Matsumoto, 1992; van SO42− depletion is controlled primarily by AOM at MD10-3262, MD10-
Lith et al., 2003). As magnesite is considered to be either an evaporite 3291 and MD10-3293 coring sites, and partly controlled by HSR at sites
mineral precipitated from concentrated seawater or an alteration- MD10-3274, MD10-3287, MD10-3288 and MD10-3290. Calculated IAP
weathering product of mafic-ultramafic igneous rocks, it is unusual to indicates the possibility of authigenic carbonate precipitation at the
observe magnesite instead of dolomite in normal anoxic marine sedi- SMTZ. Net changes of DIC fluxes (ΔDIC-Prod) were applied to determine
ments (Muller et al., 1972; Pueyo and Urpinell, 1987; Matsumoto, the rate of DIC production from two SO42− reduction reactions and to
1992). Future studies are needed to understand the formation of derive more reliable SO42− fluxes that are representative of CH4 fluxes.

180
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Fig. 7. I− versus Br− and concentrations in sediment porewater (A) this study, (B) Site 2 (Muramatsu et al., 2007, open circles), (C) Site 3 (Muramatsu et al., 2007, open circles) and (D)
Site 1230 (Fehn et al., 2007). I−/Br− display linear correlations, similar slopes, and approximate initial I−/Br− ratios in (A) and (D). This suggests that I− and Br− concentrations are
mainly controlled by biodegradation with insignificant fluid flow. I−/Br− distributions fall in two domains ((B) and (C)), mainly controlled by biodegradation or mixing with other fluids.

These calculated CH4 fluxes are higher than other gas hydrate and Appendix A. Supplementary material
upwelling areas, with higher values in the accretionary wedge than in
the passive continental margin, offshore SW Taiwan. Such high CH4 Supplementary data associated with this article can be found, in the
flux values at an active margin indicate a broad distribution of me- online version, at http://dx.doi.org/10.1016/j.jseaes.2017.07.002.
thane-enriched sources at depth, and the fold-and-thrust structures of
the active margin supplied pathways for methane to migrate upwards. References
As relatively conservative ions, Cl−, Br− and I− concentrations can
reflect the biodegradation of sedimentary organic matter and the ex- Atkins, P.W., De Paula, J., 2006. Atkins' Physical Chemistry. Oxford University Press, New
istence of other fluids. Although constant Cl− concentrations York, Oxford, pp. 1064.
Barnes, R.O., Goldberg, E.D., 1976. Methane production and consumption in anoxic
throughout ∼30 m sediment depth suggest no influence of the gas marine sediments. Geology 4 (5), 297–300.
hydrate within the core, gas hydrates may still contribute to the high Boetius, A., Ravenschlag, K., Schubert, C.J., Rickert, D., Widdel, F., Gieseke, A., Amann,
CH4 flux, since the base of gas hydrate stability zones are about R., Jorgensen, B.B., Witte, U., Pfannkuche, O., 2000. A marine microbial consortium
apparently mediating anaerobic oxidation of methane. Nature 407, 623–626.
50–700 m below the seafloor in this study area (Chen et al., 2012). Br− Borowski, W.S., Paull, C.K., Ussler, W., 1996. Marine porewater sulfate profiles indicate
and I− concentrations increase with sediment depth, and Br− vs. I− in situ methane flux from underlying gas hydrate. Geology 24, 655–658.
concentrations display a linear correlation, similar slopes, and con- Borowski, W.S., 2004. A review of methane and gas hydrates in the dynamic, stratified
system of the Blake Ridge region, offshore southeastern North America. Chem. Geol.
sistent initial I−/Br− values. In addition to SO42− which decreases with 205, 311–346.
sediment depth at a constant rate, this result suggests diffusion as the Boudreau, B.P., 1997. Diagenetic Models and Their Implementation: Modelling Transport
main transport mechanism with insignificant fluid mixing. and Reactions in Aquatic Sediments. Springer-Verlag, Berlin, pp. 414.
Buffett, B., Archer, D., 2004. Global inventory of methane clathrate: sensitivity to changes
in the deep ocean. Earth Planet. Sci. Lett. 227, 185–199.
Chatterjee, S., Dickens, G.R., Bhatnagar, G., Chapman, W.G., Dugan, B., Snyder, G.T.,
Acknowledgements Hirasaki, G.J., 2011. Porewater sulfate, alkalinity, and carbon isotope profiles in
shallow sediment above marine gas hydrate systems: a numerical modeling per-
spective. J. Geophys. Res.: Solid Earth 116, B09103.
We thank the captain and crew of the R/V Marion Dufresne and all of Chen, J.C., 2011. The Relationship Between Physical Properties of Cored Sediments and
the MD178 cruise participants for their help. We also thank the Central Gas Hydrate Accumulations in the Gas Hydrate Potential Area Offshore of
Geological Survey of Taiwan, R.O.C. (Grant No. 101-5226904000-06- Southwestern Taiwan (4/4). Report of Central Geological Survey 2011, 100-25-F, pp.
122. (in Chinese with English abstract).
01) for their generous support of this research. Chen, L.W., Chi, W.C., Liu, C.S., Shyu, C.T., Wang, Y., Lu, C.Y., 2012. Deriving regional
vertical fluid migration rates offshore southwestern Taiwan using bottom-simulating
reflectors. Mar. Geophys. Res. 33 (4), 379–388.

181
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

Chen, L.W., Chi, W.C., Wu, S.K., Liu, C.S., Shyu, C.T., Wang, Y.S., Lu, C.Y., 2014a. Two Mar. Pet. Geol. 28, 1829–1837.
dimensional fluid flow models at two gas hydrate sites offshore southwestern Taiwan. Lin, S., Hsieh, W.C., Lim, Y.C., Yang, T.F., Liu, C.S., Wang, Y., 2006. Methane migration
J. Asian Earth Sci. 92 (SI), 245–253. and its influence on sulfate reduction in the good weather Ridge region, South China
Chen, S.C., Hsu, S.K., Wang, Y., Chung, S.H., Chen, P.C., Tsai, C.H., Liu, C.S., Lin, H.S., Sea continental margin sediments. Terr. Atmos. Ocean. Sci. 17 (4), 883–902.
Lee, Y.W., 2014b. Distribution and characters of the mud diapirs and mud volcanoes Lin, A.T., Liu, C.S., Lin, C.C., Schnürle, P., Chen, G.Y., Liao, W.Z., Teng, L.S., Chuang,
off southwest Taiwan. J. Asian Earth Sci. 92, 201–214. H.R., Wu, M.S., 2008. Tectonic features associated with the overriding of an accre-
Chi, W.C., Reed, D.L., Liu, C.S., Lunberg, N., 1998. Distribution of the bottom simulating tionary wedge on top of a rifted continental margin: an example from Taiwan. Mar.
reflector in the offshore Taiwan collision zone. Terr. Atmos. Ocean. Sci. 9, 779–793. Pet. Geol. 255, 186–203.
Chow, J., Lee, J.S., Sun, R., Liu, C.S., Lundberg, N., 2000. Characteristics of the bottom Lin, C.C., Lin, A.T., Liu, C.S., Chen, G.Y., Liao, W.Z., Schnürle, P., 2009. Geological
simulating reflectors near mud diapirs: offshore southwestern Taiwan. Geo-Mar. Lett. controls on BSR occurrences in the incipient arc-continent collision zone off south-
20 (1), 3–9. west Taiwan. Mar. Pet. Geol. 26 (7), 1118–1131.
Chuang, P.C., Yang, T.F., Lin, S., Lee, H.F., Lan, T.F., Hong, W.L., Liu, C.S., Chen, J.C., Liu, C.S., Huang, I.L., Teng, L.S., 1997. Structural features off southwestern Taiwan. Mar.
Wang, Y., 2006. Extremely high methane concentration in bottom water and cored Geol. 137, 305–319.
sediments from offshore southwestern Taiwan. Terr. Atmos. Ocean. Sci. 17 (4), Liu, C.S., Liu, S.Y., Lallemand, S., Lundberg, N., Reed, D.L., 1998. Digital elevation model
903–920. offshore Taiwan and its tectonic implication. Terr. Atmos. Ocean. Sci. 9, 705–738.
Chuang, P.C., Yang, T.F., Hong, W.L., Lin, S., Sun, C.H., Lin, A.T.S., Chen, J.C., Wang, W., Liu, C.S., Deffontaines, B., Lu, C.Y., Lallemand, S., 2004. Deformation patterns of an ac-
Chung, S.H., 2010. Estimation of methane flux offshore SW Taiwan and the influence cretionary wedge in the transition zone from subduction to collision offshore
of tectonics on gas hydrate accumulation. Geofluids 10 (4), 1–14. southwestern Taiwan. Mar. Geophys. Res. 25 (1–2), 123–137.
Chuang, P.C., Dale, A.W., Wallmann, K., Haeckel, M., Yang, T.F., Chen, N.C., Chen, H.C., Liu, C.S., Schnürle, P., Wang, Y., Chung, S.H., Chen, S.C., Hsiuan, T.H., 2006. Distribution
Cheng, H.W., Lin, S., Sun, C.H., You, C.F., Hong, C.S., Wang, Y., Chung, S.H., 2013. and characters of gas hydrate offshore of southwestern Taiwan. Terr. Atmos. Ocean.
Relating sulfate and methane dynamics to geology: accretionary prism offshore SW Sci. 17, 615–644.
Taiwan. Geochem. Geophys. Geosyst. 14 (7), 2523–2545. Lu, Z., Hensen, C., Fehn, U., Wallmann, K., 2008. Halogen and 129I systematics in gas
Claypool, G.E., Milkov, A.V., Lee, Y.J., Torres, M.E., Borowski, W.S., Tomaru, H., 2006. hydrate fields at the northern Cascadia margin (IODP Expedition 311): insights from
Microbial methane generation and gas transport in shallow sediments of an accre- numerical modeling. Geochem. Geophys. Geosyst. 9, Q10006.
tionary complex, southern Hydrate Ridge (ODP Leg 204), offshore Oregon, USA. Luo, M., Dale, A.W., Haffert, L., Haeckel, M., Koch, S., Crutchley, G., De Stigter, H., Chen,
Proc. ODP Sci. Results 204, 1–52. D., Greinert, J., 2016. A quantitative assessment of methane cycling in Hikurangi
Cragg, B.A., Parkes, R.J., Fry, J.C., Weightman, A.J., Rochelle, P.A., Maxwell, J.R., 1996. Margin sediments (New Zealand) using geophysical imaging and biogeochemical
Bacterial populations and processes in sediments containing gas hydrates (ODP Leg modeling. Geochem. Geophys. Geosyst. 17, 4817–4835.
146: Cascadia Margin). Earth Planet. Sci. Lett. 139, 497–507. Martin, J.B., Gieskes, J.M., Torres, M., Kastner, M., 1993. Bromine and iodine in Peru
Dale, A.W., Somrner, S., Haeckel, M., Wallmann, K., Linke, P., Wegener, G., Pfamkudx, margin sediments and pore fluids: implications for fluid origins. Geochim.
O., 2010. Pathways and regulation of carbon, sulfur and energy transfer in marine Cosmochim. Acta 51, 4377–4389.
sediments overlying methane gas hydrates on the Opouawe Bank (New Zealand). Matsumoto, R., 1992. Diagenetic dolomite, calcite, rhodochrosite, magnesite, and lans-
Geochim. Cosmochim. Acta 74, 5763–5784. fordite from Site 799, Japan Sea – implications for depositional environments and the
Dickens, G.R., 2001. Sulfate profiles and barium fronts in sediment on the Blake Ridge: diagenesis of organic-rich sediments. In: Pisciotto, K.A., Ingle Jr.J.C., von Breymann,
present and past methane fluxes through a large gas hydrate reservoir. Geochim. M.T., Barron, J. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results,
Cosmochim. Acta 65, 529–543. 127/128 (Pt. 1): College Station, TX (Ocean Drilling Program), pp. 75–98.
Dickson, A.G., Goyet, C., 1994. Handbook of Methods for the Analysis of the Various Mazzini, A., Svensen, H., Hovland, M., Planke, S., 2006. Comparison and implications
Parameters of the Carbon Dioxide System in Sea Water. CO2 Science Team Report from strikingly different authigenic carbonates in a Nyegga complex pockmark, G11,
2.0. U.S. Department of Energy, Washington, D.C. Norwegian Sea. Mar. Geol. 231, 89–102.
Egeberg, P.K., Dickens, G.R., 1999. Thermodynamic and porewater halogen constraints Meister, P., McKenzie, J.A., Vasconcelos, C., Bernasconi, S., Frank, M., Gutjahr, M.,
on gas hydrate distribution at ODP Site 997 (Blake Ridge). Chem. Geol. 153, 53–79. Schrag, D.P., 2007. Dolomite formation in the dynamic deep biosphere: results from
Fehn, U., Lu, Z., Tomaru, H., 2006. Data report: 129I/I ratios and halogen concentrations the Peru Margin. Sedimentology 54, 1007–1031.
in porewaters of the Hydrate Ridge and their relevance for the origin of gas hydrates: Milkov, A.V., 2004. Global estimates of hydrate-bound gas in marine sediments: how
a progress report. In: Proceedings of ODP, Sciences Results 204, 1–25, MS 204SR-107. much is really out there? Earth Sci. Rev. 66 (3–4), 183–197.
Fehn, U., Snyder, G.T., Muramatsu, Y., 2007. Iodine as a tracer of organic material: 129I Moore, T.S., Murray, R.W., Kurtz, A.C., Schrag, D.P., 2004. Anaerobic methane oxidation
results from gas hydrate systems and fore arc fluids. J. Geochem. Explor. 95, 66–80. and the formation of dolomite. Earth Planet. Sci. Lett. 229, 141–154.
Greinert, J., Gerhard, B., Suess, E., 2001. Gas hydrate-associated carbonates and methane Muller, G., Irion, G., Forstner, U., 1972. Formation and diagenesis of inorganic Ca-Mg
venting at hydrate ridge: Classification, distribution, and origin of authigenic carbonates in the lacustrine environment. Sci. Nat. 59 (4), 158–164.
lithologies. In: In: Paull, C.K., Dillon, W.P. (Eds.), Natural gas hydrates: Occurrence, Muramatsu, Y., Doi, T., Tomaru, H., Fehn, U., Takeuchi, R., Matsumoto, R., 2007.
distribution, and detection, vol. 124. American Geophysical Union Geophysical Halogen concentrations in porewaters and sediments of the Nankai Trough, Japan:
Monograph, pp. 99–113. implications for the origin of gas hydrates. Appl. Geochem. 22, 534–556.
Hiruta, A., Snyder, G.T., Tomaru, H., Matsumoto, R., 2009. Geochemical constraints for Niewöhner, C., Hensen, C., Kasten, S., Zabel, M., Schulz, H.D., 1998. Deep sulfate re-
the formation and dissociation of gas hydrate in an area of high methane flux, eastern duction completely mediated by anaerobic methane oxidation in sediments of the
margin of the Japan Sea. Earth Planet. Sci. Lett. 279 (3–4), 326–339. upwelling area off Namibia. Geochim. Cosmochim. Acta 62, 455–464.
Huang, C.Y., Wu, W.Y., Chang, C.P., Tsao, S., Yuan, P.B., Lin, C.W., Xia, K.Y., 1997. Orcutt, B.N., Sylvan, J.B., Knab, N.J., Edwards, K.J., 2011. Microbial ecology of the dark
Tectonic evolution of accretionary prism in the arc-continent collision terrane of ocean above at and below the seafloor. Microbiol. Mol. Biol. Rev. 75, 361–422.
Taiwan. Tectonophysics 281, 31–51. Pueyo, M.J.J., Urpinell, I.M., 1987. Magnesite formation in recent playa lakes, Los
Jiang, W.T., Chen, J.C., Huang, B.J., Chen, C.J., Lee, Y.T., Huang, P.R., Lung, C.C., Huang, Monegros, Spain. In: In: Marshall, J.D. (Ed.), Diagenesis of Sedimentary Sequences,
S.W., 2006. Mineralogy and physical properties of cored sediments from the gas vol. 36. Geological Society London Special Publications, pp. 119–122.
hydrate potential area of offshore Southwestern Taiwan. Terr. Atmos. Ocean. Sci. 17 Reed, D.L., Lundberg, N., Liu, C.S., Kuo, B.Y., 1992. Structural relations along the margins
(4), 981–1007. of the offshore Taiwan accretionary wedge: implications for accretion and crustal
Jiang, W.T., 2011. Clay and Authigenic Minerals Studies in the Cored Sediments from the kinematics. ACTA Geol. Taiwan. Sci. Rep. Nat. Taiwan Univ. 30, 105–122.
Gas Hydrate Potential Area Offshore of Southwestern Taiwan (4/4). Report of Central Regnier, P., Dale, A.W., Arndt, M., LaRowe, D.E., Mogollón, J., Van Cappellen, P., 2011.
Geological Survey 2011, 100-25-D, pp. 102. (in Chinese with English abstract). Quantitative analysis of anaerobic oxidation of methane (AOM) in marine sediments:
Jiang, W.T., 2015. Electron Microscope Analyses of Authigenic minerals in the cored a modeling perspective. Earth Sci. Rev. 106 (1–2), 105–130.
sediments from the Gas Hydrate Potential Area of Offshore Southwestern Taiwan (4/ Schnurle, P., Hsiuan, T.-H., Liu, C.-S., 1999. Constraints on free gas and gas hydrate
4). Report of Central Geological Survey 2015, 104-11-G, pp. 121. (in Chinese with bearing sediments from multi-channel seismic data, offshore southwestern Taiwan.
English abstract). Pet. Geol. Taiwan 33, 21–42.
Joye, S.B., Boetius, A., Orcutt, B.N., Montoya, J.P., Schulz, H.N., Erickson, M.J., Lugo, Schnürlea, P., Liu, C.S., Lin, A.T., Lin, S., 2011. Structural controls on the formation of
S.K., 2004. The anaerobic oxidation of methane and sulfate reduction in sediments BSR over a diapiric anticline from a dense MCS survey offshore southwestern Taiwan.
from Gulf of Mexico cold seeps. Chem. Geol. 205, 219–238. Mar. Pet. Geol. 28 (10), 1932–1942.
Kastner, M., Claypool, G., Robertson, G., 2008. Geochemical constraints on the origin of Schulz, H.D., 2006. Conceptual models and computer models. In: Schulz, H.D., Zabel, M.
the pore fluids and gas hydrate distribution at Atwater Valley and Keathley Canyon, (Eds.), Marine Geochemistry, 2. Edn. Springer-Verlag, Berlin, Heidelberg, New York,
northern Gulf of Mexico. Mar. Pet. Geol. 25, 860–872. pp. 513–547.
Kelts, K., McKenzie, J.A., 1982. Diagenetic dolomite formation in quaternary anoxic Siegel, F.R., 1975. Applied Geochemistry. Wiley, New York, pp. 353.
diatomaceous muds of deep-sea drilling project Leg-64, Gulf of California. Deep Sea Snyder, G.T., Hiruta, A., Matsumoto, R., Dickens, G.R., Tomaru, H., Takeuchi, R.,
Drill. Project Initial Rep. 64, 553–569. Komatsubara, J., Ishida, Y., Yu, H., 2007. Pore water profiles and authigenic mi-
Komada, T., Burdige, J.D., Li, H.-L., Magen, C., Chanton, P.J., Cada, A.K., 2016. Organic neralization in shallow marine sediments above the methane-charged system on
matter cycling across the sulfate-methane transition zone of the Santa Barbara Basin, Umitaka Spur, Japan Sea. Deep Sea Res. Part II: Topical Stud. Oceanogr. 54,
California Borderland. Geochim. Cosmochim. Acta 176, 259–278. 1216–1239.
Kvenvolden, K.A., 1993. Gas hydrates-geological perspective and global change. Rev. Suess, E., Carson, B., Ritger, S.D., Moore, J.C., Jones, M.L., Kulm, L.D., Cochrane, G.R.,
Geophys. 31, 173–187. 1985. Biological communities at vent sites along the subduction zone off Oregon.
Li, T.H., Gregory, S., 1974. Diffusion of ions in seawater and deep-sea sediments. Bull. Biol. Soc. Wash. 6, 475–484.
Geochim. Cosmochim. Acta 38, 703–714. Suess, E., Torres, M.E., Bohrmann, G., Collier, R.W., Rickert, D., Goldfinger, C., Linke, P.,
Lim, Y.C., Lin, S., Yang, T.F., Chen, Y.G., Liu, C.S., 2011. Variations of methane induced Heuser, A., Sahling, H., Heeschen, K., Jung, C., Nakamura, K., Greinert, J.,
pyrite formation in the accretionary wedge sediments offshore southwestern Taiwan. Pfannkuche, O., Trehu, A., Klinkhammer, G., Whiticar, M.J., Eisenhauer, A., Teichert,

182
C.-Y. Hu et al. Journal of Asian Earth Sciences 149 (2017) 172–183

B., Elvert, M., 2001. Sea floor methane hydrates at Hydrate Ridge, Cascadia Margin. bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation.
In: In: Paull, C., Dillon, W. (Eds.), Natural Gas Hydrates: Occurrence, Distribution, Geobiology 1, 71–79.
and Detection, vol. 124. American Geophysical Union, Monograph Series, pp. 87–98. Vasconcelos, C., McKenzie, J.A., Bernasconi, S., Grujic, D., Tien, A.J., 1995. Microbial
Sun, C.H., Chang, S.C., Kuo, C.L., Wu, J.C., Shao, P.H., Oung, J.N., 2010. Origins of mediation as a possible mechanism for natural dolomite formation at low tempera-
Taiwan’s mud volcanoes: Evidence from geochemistry. J. Asian Earth Sci. 37 (2), tures. Nature 377, 220–222.
105–116. Wallmann, K., Drews, M., Aloisi, G., Bohrmann, G., 2006a. Methane discharge into the
Teng, L.S., 1990. Geotectonic evolution of late Cenozoic arc-continent collision in Black Sea and the global ocean via fluid flow through submarine mud volcanoes.
Taiwan. Tectonophysics 183, 57–76. Earth Planet. Sci. Lett. 248, 544–559.
Torres, M.E., McManus, J., Hammond, D.E., de Angelis, M.A., Heeschen, K.U., Colbert, Wallmann, K., Aloisi, G., Haeckel, M., Obzhirov, A., Pavlova, G., Tishchenko, P., 2006b.
S.L., Tryon, M.D., Brown, K.M., Suess, E., 2002. Fluid and chemical fluxes in and out Kinetics of organic matter degradation, microbial methane generation, and gas hy-
of sediments hosting methane hydrate deposits on Hydrate Ridge, OR: I. Hydrological drate formation in anoxic marine sediments. Geochim. Cosmochim. Acta 70 (15),
provinces. Earth Planet. Sci. Lett. 201, 525–540. 3905–3927.
Torres, M.E., Wallmann, K., Trehu, A.M., Bohrmann, G., Borowski, W.S., Tomaru, H., Warthmann, R., van Lith, Y., Vasconcelos, C., McKenzie, J.A., Karpoff, A., 2000.
2004a. Gas hydrate growth, methane transport, and chloride enrichment at the Bacterially induced dolomite precipitation in anoxic culture experiments. Geology
southern summit of Hydrate Ridge, Cascadia margin off Oregon. Earth Planet. Sci. 28, 1091–1094.
Lett. 226, 225–241. Wehrmann, L.M., Risgaard-Petersen, N., Schrum, H.N., Walsh, E.A., Huh, Y., Ikehara, M.,
Torres, M.E., Teichert, B.M.A., Trehu, A.M., Borowski, W., Tomaru, H., 2004"/>b. Pierre, C., D'Hondt, S., Ferdelman, T.G., Ravelo, A.C., Takahashi, K., Zarikian, C.A.,
Relationship of porewater freshening to accretionary processes in the Cascadia 2011. Coupled organic and inorganic carbon cycling in the deep subseafloor sediment
margin: fluid sources and gas hydrate abundance. Geophys. Res. Lett. 31, L22305. of the northeastern Bering Sea Slope (IODP Exp. 323). The Integrated Ocean Drilling
Treude, T., Boetius, A., Knittel, K., Wahanb, K., Jargensen, B.B., 2003. Anaerobic oxi- Program Expedition 323 Scientific Party. Chem. Geol. 284, 251–261.
dation of methane above gas hydrates at Hydmte Ridge, NE Pacific Ocean. Mar. Ecol. Yang, T.F., Chuang, P.C., Lin, S., Chen, J.C., Wang, Y., Chung, S.H., 2006. Methane
Prog. Ser. 264, 1–14. venting in gas hydrate potential area offshore of SW Taiwan: evidence of gas analysis
Ussler, W., Paull, C.K., 2008. Rates of anaerobic oxidation of methane and authigenic of water column samples. Terr. Atmos. Ocean. Sci. 17 (4), 933–950.
carbonate mineralization in methane-rich deep-sea sediments inferred from models Yu, H.S., Song, G.S., 2000. Submarine physiographic features in Taiwan Region and their
and geochemical profiles. Earth Planet. Sci. Lett. 266, 271–287. geological significance. J. Geol. Soc. China 43 (2), 267–286.
van Lith, Y., Warthmann, R., Vasconcelos, C., McKenzie, J.A., 2003. Sulphate-reducing

183

You might also like