Evidence for Microbial Carbon and Sulfur Cycling in Please share

advertisement
Evidence for Microbial Carbon and Sulfur Cycling in
Deeply Buried Ridge Flank Basalt
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Lever, M. A., O. Rouxel, J. C. Alt, N. Shimizu, S. Ono, R. M.
Coggon, W. C. Shanks, et al. “Evidence for Microbial Carbon and
Sulfur Cycling in Deeply Buried Ridge Flank Basalt.” Science
339, no. 6125 (March 15, 2013): 1305–1308.
As Published
http://doi.org/10.1126/science.1229240
Publisher
American Association for the Advancement of Science (AAAS)
Version
Original manuscript
Accessed
Wed May 25 23:38:39 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/87684
Terms of Use
Creative Commons Attribution-Noncommercial-Share Alike
Detailed Terms
http://creativecommons.org/licenses/by-nc-sa/4.0/
1
1
Evidence for Microbial Carbon and Sulfur Cycling in Deeply
2
Buried Ridge Flank Basalt
3
4
Mark A. Lever1,2*, Olivier Rouxel3,4, Jeffrey C. Alt5, Nobumichi Shimizu3, Shuhei
5
Ono6, Rosalind M. Coggon7, Wayne C. Shanks, III8, Laura Lapham2,9, Marcus
6
Elvert10, Xavier Prieto-Mollar10, Kai-Uwe Hinrichs10, Fumio Inagaki11, and Andreas
7
Teske1*
8
9
1
Department of Marine Sciences, University of North Carolina, Chapel Hill, NC
10
27516, USA.
11
2
12
8000 Aarhus C, Denmark.
13
3
Woods Hole Oceanographic Institution, Woods Hole, MA 02543, USA.
14
4
IFREMER, Centre de Brest, 29280 Plouzané, France.
15
5
Department of Earth and Environmental Sciences, University of Michigan, Ann
16
Arbor, MI 48109, USA.
17
6
18
of Technology, Cambridge, MA 02139, USA.
19
7
20
Kensington Campus, London SW7 2AZ, UK.
21
8
U.S. Geological Survey, Denver, CO 80225, USA.
22
9
Present address: University of Maryland Center for Environmental Science,
23
Solomons, MD 20688, USA.
Center for Geomicrobiology, Department of BioScience, Aarhus University, DK-
Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute
Department of Earth Science and Engineering, Imperial College London, South
2
24
10
25
Marine Environmental Sciences, University of Bremen, D-28334 Bremen, Germany.
26
11
27
for Marine-Earth Science and Technology, Nankoku, Kochi 783-8502, Japan.
28
*Corresponding authors, email: mark.lever@biology.au.dk, teske@email.unc.edu.
Organic Geochemistry Group, Department of Geosciences & MARUM Center for
Geomicrobiology Group, Kochi Institute for Core Sample Research, Japan Agency
29
30
31
One sentence summary:
32
Functional gene, geochemical, and incubation evidence reveal active, methane-
33
and sulfur-cycling microbial communities in deep basaltic ocean crust.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
3
50
Sediment-covered basalt on ridge flanks constitutes most of Earth’s oceanic
51
crust, but the composition and metabolic function of its microbial ecosystem is
52
largely unknown. Here, we demonstrate the presence of methane- and sulfur (S)-
53
cycling microbes in subseafloor basalt of the eastern flank of the Juan de Fuca
54
Ridge. Depth horizons with functional genes indicative of methane-cycling and
55
sulfate-reducing microorganisms are enriched in solid-phase S and total organic
56
carbon (TOC), host δ 13C- and δ 34S-isotopic values with a biological imprint, and
57
show clear signs of microbial activity when incubated in the laboratory.
58
Downcore changes in C and S cycling show discrete geochemical intervals with
59
chemoautotrophic δ 13C-TOC signatures potentially attenuated by heterotrophic
60
metabolism.
61
Subseafloor basaltic crust represents the largest habitable zone by volume on
62
Earth (1). Chemical reactions of basalt with seawater flowing through fractures
63
(veins) release energy that may support chemosynthetic communities. Microbes
64
exploiting these reactions are known from basalt exposed at the seafloor, where the
65
oxidation of reduced S and iron (Fe) from basalt with dissolved oxygen and nitrate
66
from seawater supports high microbial biomass and diversity (2, 3). Multiple lines of
67
evidence that include textural alterations (4), depletions in δ34S-pyrite (FeS2) (5) and
68
δ13C-dissolved inorganic carbon (DIC) (6), and DNA sequences from borehole
69
observatories (7, 8) suggest active microbial communities in subseafloor basalt. Yet,
70
direct evidence of these communities is still missing. In this study we combine
71
sequencing of genes diagnostic of microbial methane- and S-cycling with
72
geochemical and isotopic analyses of C- and S-pools and laboratory-based
73
incubations to identify microbial ecosystem components in deep subseafloor basalt.
74
The 3.5 million-year-old basement at Hole U1301B was sampled during
75
Integrated Ocean Drilling Program (IODP) Expedition 301 in 2004 (Fig. S1) (9). Site
76
U1301 is covered by a 265-m thick sediment layer and lies ~2 km south of ODP Site
4
77
1026, which it resembles in temperature profile, lithology, and sediment chemistry
78
(9). Given anticipated poor recovery due to brecciation of the upper basement (265-
79
350 meters below seafloor; mbsf), coring was restricted to an interval of pillow
80
basalts and massive lavas (351-583 mbsf). Sulfate concentrations (~16mM) and vein
81
carbonates indicate basalt fluids are derived from seawater, which enters ~80 km
82
south at Grizzly Bare outcrop and discharges near U1301B, at Baby Bare and Mama
83
Bare outcrops (9, 10; Fig. S1B). Yet, the basement at U1301 differs from seafloor-
84
exposed basalt in its uniformly high temperature (~64°C) (9) and lack of fresh
85
photosynthesis-derived organic matter, dissolved oxygen and nitrate (7, 11). These
86
conditions preclude chemosynthetic S- and Fe-oxidation reactions that require oxygen
87
or nitrate as electron acceptors (12), but may enable growth of strict anaerobes, such
88
as many methanogens, methane-oxidizing archaea and sulfate reducers.
89
By sequencing genes encoding the α subunit of methyl coenzyme M reductase
90
(mcrA), a gene unique to methanogens and anaerobic methane-oxidizers (13), and the
91
β subunit of dissimilatory sulfite reductase (dsrB), a gene found in sulfate- and
92
sulfite-reducing microbes (14), we demonstrate the presence of key genes of methane-
93
cycling and sulfate-reducing microbes (Methods in Supporting Online Material). We
94
detect mcrA in 5 of the 10 samples and dsrB in 4 of the 6 samples tested (Table S1).
95
The phylogenetic diversity of mcrA genes is restricted to two groups: the Juan
96
de Fuca Methanogen Group (JdFMG), which falls into an uncultivated cluster within
97
the Methanosarcinales, and anaerobic methane-oxidizing archaea (ANME-1; Fig.
98
1A). Close relatives of the JdFMG were first reported as ‘Unidentified Rice Field Soil
99
mcrA’ from paddy soil (15), and have since also been found in marine habitats,
100
including JdF Ridge hydrothermal vent chimneys and seafloor-exposed basalt ~100
101
km west of U1301B (Fig. S2) (16-17). Recently, a close relative of JdFMG, which
102
uses hydrogen (H2), acetate, methanol and trimethylamine as energy substrates for
5
103
methanogenesis, was enriched in wetland soil (18). Given that substrate use follows
104
phylogenetic divisions in mcrA (19), the JdFMG are likely to be methanogens with
105
similar substrate requirements. ANME-1 occur widely in marine sediments and
106
methane seeps, and are believed to gain energy from the anaerobic oxidation of
107
methane (AOM) (20). Two distinct ANME phylotypes occur at U1301, one closely
108
related to ANME-1 from methane seeps, and another clustering with only one other
109
sequence, from subseafloor sediment (Fig. S3). We detect JdFMG in four, and
110
ANME-1 in three out of ten basalt samples. Two samples contain both groups (Table
111
S1).
112
The phylogenetic diversity of dsrB is limited to one group, the Juan de Fuca
113
Sulfate Reducing Group (JdFSRG), which falls into Cluster IV, a deeply-branching
114
dsrB cluster without cultured members, first reported from hydrothermal sediment
115
(Fig. 1B and S4, Table S1) (21). Remarkably, the only other dsr sequences reported
116
so far from the subseafloor, i.e. sediment of the Peru Margin (22), fall into this
117
cluster, which is widespread in shallow marine sediment and terrestrial aquifers.
118
The study of solid-phase S-pools through analyses of acid-volatile sulfide
119
(AVS), chromium-reducible S (CRS), and sulfate-S (SO4-S) provides insights to
120
redox processes (5, 23), and shows a relationship with dsrB distributions at U1301:
121
dsrB was only found in a relatively reduced ‘intermediate depth interval’ (~430-520
122
mbsf, 14R-26R), in samples with AVS as the main S pool in alteration halos (14R-1-
123
11) or host rock (17R-170, 20R-1-57, 23R-2-21; Fig. S5, Table S1). Samples from
124
this interval have higher AVS, CRS, and total S (Fig. S5, Table S2), contain large
125
pyrite fronts (14R-1-65P, 15R-4-142P; Fig. S5), and have lower δ34S-AVS, -CRS, and
126
-SO4-S, compared to the more oxidized upper (1R-12R) and lower coring intervals
127
(30R-36R; Fig. S6, Table S1). Consistent with higher Fe3+/FeTotal ratios, which
128
indicate halos to be more oxidized than host rock (Table S1), pyrite is generally
6
129
absent from halos or veins. Outside the intermediate depth interval, the near absence
130
of pyrite from host rock, and mixed clay-Fe-oxyhydroxide-dominated halos and veins
131
are further evidence of pervasive oxidative alteration.
132
To examine micro- and macroscale variations in microbial S-cycling, the δ34S
133
of pyrite grains were analyzed (Tables S1 and S3, Fig. 2). Though variable, the δ34S-
134
pyrite grains (-72.4 to 1.2‰; Table S3) are typically lower than those of AVS (-9.3 to
135
-0.2‰), CRS (-13.7 to 0‰), SO4-S (-6.5 to 0‰), mantle S (0‰) (5), dissolved sulfate
136
in bottom sediments at ODP Site 1026 (+30‰) (24), or seawater (+21‰; Fig. 2).
137
Locally, the δ34S of pyrite grains reach very negative values (-72‰), consistent with
138
the addition of highly 34S-depleted secondary sulfide to basement rock (23). These
139
low δ34S-pyrite values indicate single-step sulfate reduction (25) or repeated cycles of
140
sulfate reduction and S oxidation (26). The co-occurrence of low δ34S-pyrite, dsrB,
141
and mcrA of ANME-1 in two samples (14R-1-11, 17R-1-70) suggests local coupling
142
between methane and S-cycling by sulfate-dependent AOM.
143
Depth profiles of TOC content, δ13C-TOC, and δ13C-carbonate at U1301B are
144
consistent with functional gene- and 34S-data (Fig. 3). The TOC content is highest in
145
the intermediate depth interval in cores with mcrA, dsrB, and low δ34S-pyrite (Fig.
146
3A, Table S4). The δ13C-TOC is in the range of dissolved organic C (DOC) in fluids
147
from nearby 1026B and Baby Bare Springs (BBS; Fig. 3B, Table S4) and lower than
148
seawater DOC (-21.1‰; 6). The δ13C-carbonate is higher than δ13C-DIC at 1026B or
149
BBS (Fig. 3C, Table S5) and overlaps with δ13C-DIC of bottom seawater (-1.4‰; 10).
150
δ13C-TOC values in the upper coring interval (4R-5R) and near the bottom
151
(23R-26R; -34.6 to -32.0‰) are close to δ13C-DOC from nearby BBS (-34.6‰; Fig.
152
3B). The absence of O2 and the high 13C-TOC depletion relative to carbonate (~ -30
153
to -35‰) suggest C fixation by the reductive acetyl CoA pathway – an anaerobic
7
154
pathway found in all methanogens and acetogens, and certain sulfate and iron
155
reducers (Fig. S7, Tables S6 and S7; 27). The presence of dsrB but not mcrA in these
156
samples suggests that sulfate reducers or other groups, but not methanogens, produce
157
this low δ 13C-TOC.
158
δ 13C-TOC at the top (2R) and in the intermediate depth interval (-28.4 to -
159
21.6‰) are close to δ13C-DOC from borehole 1026B (-26.1‰, Fig. 3B; 6). The 13C-
160
depletion relative to inorganic C is lower than in the other layers (~ -20 to -26‰), but
161
also falls in the range of the reductive acetyl CoA pathway (Table S7), and, consistent
162
with mcrA detection, could be impacted by autotrophic methanogenesis. In addition,
163
elevated heterotrophic activity is possible, since degradation of chemoautotrophy-
164
derived OC, e.g. by AOM, methanogenesis or fermentation, would lower the δ13C-
165
carbonate and potentially raise the δ13C-TOC. The alternative explanation, enhanced
166
breakdown of photosynthesis-derived OC in the intermediate depth interval, is
167
unlikely given that sediment inclusions are absent (9). Similarly, influx of labile DOC
168
or unaltered DIC from seawater is incompatible with the 7-11 kyr greater DOC age
169
compared to bottom seawater and the 4-8‰ decrease in δ13C-DIC along the flowpath
170
from Grizzly Bare outcrop to 1026B and BBS, respectively (6, 10).
171
172
To rule out a fossil origin of functional genes, C and S compounds in subseafloor
173
basalt, we incubated pieces from the interior of three rocks used for functional gene
174
analyses (1R-1-79, 14R-1-11, 23R-2-21) at 65°C in anoxic, sulfate-rich media
175
containing H2, acetate, methanol, and dimethyl sulfide as energy substrates (Table
176
S8). After two years, aliquots were transferred to fresh media, and incubated for
177
another five years using triple-autoclaved basalt pieces as substrata. By then, low
178
concentrations of 13C-depleted methane (-54 to -65‰) had formed indicating the
179
presence of active methanogenic microorganisms (Table S9).
8
180
The variability in δ34S-pyrite, δ13C-TOC and δ13C-carbonate indicates that
181
micro- and macro-scale geochemical changes related to mineralogy, fracturing and/or
182
fluid flow strongly influence microbial activity. These chemical microniches may
183
explain the coexistence of sulfate reducers and methanogens at U1301 and in other
184
igneous habitats, despite higher energy yields of sulfate reduction compared to
185
methanogenesis (28). In addition, some methanogens can survive in the presence of
186
sulfate reducers by consuming non-competitive methyl substrates (19); this
187
explanation is consistent with the ability of a close relative of JdFMG to use methanol
188
(18), and the production of biogenic methane in basalt incubations containing sulfate
189
and methanol (Table S9, Fig. S8). The sources of energy substrates to sulfate reducers
190
and methanogens at U1301B remain unknown. A potential source of H2 is the
191
oxidative alteration of Fe or S minerals, while short-chain organic acids and alcohols
192
might be produced by breakdown of TOC (19, 28-29) or Fischer-Tropsch-type
193
synthesis (30).
194
Samples from the eastern flank of the Juan de Fuca Ridge provide the first
195
integrated view of microbial community structure and activity in subseafloor basalt.
196
Our analyses of functional gene, geochemical, and isotopic data reveal strong micro-
197
and macroscale heterogeneity of microbial C- and S-cycling, and indicate a key role
198
of chemoautotrophy in supporting Earth’s oceanic crustal biosphere.
199
200
201
202
203
204
9
205
References
206
1. W. Bach, K. J. Edwards, Geochim. Cosmochim. Acta 67, 3871 (2003).
207
2. K. J. Edwards, T. M. McCollom, H. Konishi, P. R. Buseck. Geochim.
208
Cosmochim. Acta 67, 2843 (2003).
209
3. C. M. Santelli et al., Nature 453, 653 (2008).
210
4. M. R. Fisk, S. J. Giovannoni, I. H. Thorseth, Science 281, 978 (1998).
211
5. O. Rouxel, S. Ono, J. Alt, D. Rumble, J. Ludden, Earth Planet. Sci. Lett. 268,
212
110 (2008).
213
6. M. D. McCarthy et al., Nature Geosci 4, 32 (2011).
214
7. J. P. Cowen et al., Science 299, 120 (2003).
215
8. B. N. Orcutt et al., ISME J. 5, 692 (2011)
216
9. A. T. Fisher, T. Urabe, A. Klaus, Expedition 301 Scientists, Proc. IODP 301
217
218
219
(2005).
10. B. D. Walker, M. D. McCarthy, A. T. Fisher, T. P. Guilderson, Mar. Chem.
108, 123 (2008).
220
11. C. G. Wheat et al., Geochem. Geophys. Geosyst. 11, 1 (2010).
221
12. A. Schippers, B. B. Jørgensen, Geochim. Cosmochim. Acta 66, 85 (2002).
222
13. M. W. Friedrich, Meth. Enzymol. 397, 428 (2005).
223
14. M. Wagner, A. J. Roger, J. L. Flax, G. A. Brusseau, D. A. Stahl, Appl.
224
225
226
Environ Microbiol. 75, 7086 (1998).
15. T. Lueders, K.-J. Chin, R. Conrad, M. Friedrich, Environ. Microbiol. 3, 194
(2001).
227
16. F. Wang et al., Proc. Nat. Acad. Sci. U.S.A. 106, 4840 (2009).
228
17. O. U. Mason et al., ISME J. 3, 231 (2009).
229
18. G. Zhang et al., Environ. Microbiol. 10, 1850 (2008).
230
19. W. B. Whitman, T. L. Bowen, D. R. Boone, The Prokaryotes 3, 165 (2006).
231
20. K. Knittel, A. Boetius, Ann. Rev. Microbiol. 63, 311 (2009).
10
232
233
21. A. Dhillon, A. Teske, J. Dillon, D. A. Stahl, M. L. Sogin, Appl. Environ.
Microbiol. 69, 2765 (2003).
234
22. G. Webster et al., FEMS Microbiol. Ecol. 58, 65 (2006).
235
23. S. Ono, N.S. Keller, O. Rouxel, J. C. Alt, Geochim. Cosmochim. Acta 87, 323
236
237
238
(2012).
24. M. D. Rudnicki, H. Elderfield, B. Spiro, Geochim. Cosmochim. Acta 65, 777
(2001).
239
25. M. S. Sim, T. Bosak, S., Science 333, 74 (2011).
240
26. D. E. Canfield, B. Thamdrup, Science 266, 1973 (1994).
241
27. A. L. Zerkle, C. H. House, S. L. Brantley, Am. J. Sci. 305, 467 (2005).
242
28. H.-T. Lin, J. P. Cowen, E. J. Olson, J. P. Amend, M. D. Lilley, Geochim.
243
Cosmochim. Acta 85, 213 (2012).
244
29. M. A. Lever, et al., Geomicrobiol. J. 27, 183 (2010).
245
30. T. M. McCollom, J. S. Seewald, Chem. Rev. 107, 382 (2007).
246
31. M. A. Lever et al., Geomicrobiol. J. 23, 517 (2006).
247
32. J. F. Biddle et al., Proc. Nat. Acad. Sci. U.S.A. 103, 3846 (2006).
248
33. M. A. Lever, Ph.D. Dissertation. University of North Carolina at Chapel Hill,
249
Chapel Hill, NC (2008).
250
34. W. Ludwig et al., Nucleic Acids Res. 32, 1363 (2004).
251
35. E. M. Ripley et al., Rev. Mineral. Geochem. 73, 9 (2011).
252
36. P. Craddock, O. Rouxel, L. Ball, W. Bach, Chem. Geol. 253, 102 (2008).
253
37. A. Delacour, G. L. Früh-Green, S. M. Bernasconi, P. Schaeffer, D. S. Kelley,
254
255
256
Geochim. Cosmochim. Acta 72, 3681 (2008).
38. B. N. Popp, F. J. Sansone, T. M. Rust, D. A. Merritt, Anal. Chem. 67, 405
(1995).
257
39. C. H. House, J. W. Schopf, K. O. Stetter, Org. Geochem. 34, 345 (2003).
258
40. K. L. Londry, D. J. Des Marais, Appl. Environ. Microbiol. 69, 2942 (2003).
11
259
260
261
262
41. K. L. Londry, L. L. Jahnke, D. J. Des Marais, Appl. Environ. Microbiol. 70,
745 (2004).
42. S. Sakata, J. M. Hayes, M. Rohmer, A. B. Hooper, M. Seemann, Org.
Geochem. 39, 1725 (2008).
263
43. M. J. Alperin, T. M. Hoehler, Am. J. Sci. 309, 958 (2009).
264
44. M. Könneke, J. S. Lipp, K.-U. Hinrichs, Org. Geochem. 48, 21 (2012).
265
45. A. Gittel, K. B. Sørensen, T. L. Skovhus, K. Ingvorsen, A. Schramm, Appl.
266
267
Environ. Microbiol. 75, 7086 (2009).
46. B. O. Steinsbu et al., Int. J. Syst. Evol. Microbiol. 60, 2745 (2010).
268
269
Acknowledgments. We thank B. B. Jørgensen, M. L. Sogin and the IODP Expedition
270
301 Scientists. Funding was obtained from a Schlanger Ocean Drilling Fellowship, a
271
University of North Carolina Dissertation Completion Fellowship, a Marie-Curie
272
Intra-European Fellowship (#255135; all to M. A. Lever), the Danish National
273
Research Foundation, the Max Planck Society (both to B. B. Jørgensen), Europole
274
Mer (to O. Rouxel), the European Research Council Advanced Grant DARCLIFE (to
275
K.-U. Hinrichs), the NASA Astrobiology Institute Subsurface Biospheres (to A.
276
Teske), the Japan Society for the Promotion of Science (JSPS) Funding Program for
277
Next Generation World-Leading Researchers (NEXT Program; to F. Inagaki), and the
278
National Science Foundation (NSF-OCE 0622949 and OCE 1129631 to J. Alt, OCE-
279
0753126 to S. Ono and O. Rouxel, and NSF-ODP 0727175 and NSF-STC for Dark
280
Energy Biosphere Investigations to A. Teske). The geochemical data are fully
281
documented and available in the Supplementary Tables. The functional gene sequence
282
data are available from Genbank database.
12
283
FIGURES
284
285
286
287
Figure 1. Phylogenetic trees of functional genes. (A) McrA sequences from U1301B
288
are in bold magenta type face. Close relatives based on microarray analyses of JdF
289
Ridge hydrothermal vent chimneys and seafloor basalt are in green (16) and cyan
290
(17), respectively. (B) DsrB sequences from U1301B are in bold magenta type face,
291
and sequences from subseafloor sediment off Peru in cyan (22). Bootstrap support (in
292
%, 1,000 replications) is indicated at each branching point.
293
13
294
295
296
297
298
299
300
Figure 2. Macro-and micro-scale distribution of S-isotopic data. On the left, δ34S-
301
depth profile of pyrite granules, analyzed by laser ablation and secondary ion mass
302
spectrometry (SIMS), and bulk S pools (AVS, CRS). On the right, thin section
303
micrograph showing individual pyrite granules and their δ34S. The dashed magenta
304
line indicates the sampling depth of the thin section. The dashed black lines mark the
305
intermediate depth interval. Pyrite grains with a sufficient diameter for δ34S-
306
307
determination (10 µm) were limited to this interval. The scale bar is 200 µm.
14
308
309
310
311
312
Figure 3. Depth-related trends in (A) TOC content, (B) δ13C-TOC, and (C) δ13C-
313
carbonate. Cores with functional gene detection are indicated in A and B. Dashed
314
vertical lines indicate δ13C-DOC (B) and –DIC (C) values from 1026B and BBS.
315
Because the carbonate content of rock samples used in (A) and (B) was too low for
316
analyses, δ13C from carbonate veins are shown in (C). The reduced intermediate depth
317
interval falls between the dashed horizontal lines. All δ13C are in ‰ vs. VPDB.
Download