ArticlePDF Available

Enhanced East Pacific Rise hydrothermal activity during the last two glacial terminations

Authors:

Abstract and Figures

Searching sediment for climate signals Sediments on the ocean floor may provide clues about the interplay between ice ages and mid-ocean ridge magma production. Lund et al. present well-dated and detailed sediment records from hydrothermal activity along the East Pacific Rise. The sediments show changes in metal fluxes that are tied to the past two glaciations. Ice age changes in sea level alter magma production, which is manifested by changes in hydrothermal systems. The apparent increase in hydrothermal activity at the East Pacific Rise around the past two glacial terminations suggests some role in moderating the size of ice sheets. Science , this issue p. 478
Content may be subject to copyright.
electronic structure and tunneling probability of
the different CO molecules. We illustrate this by
calculatingthetunnelingcurrentprobabilityusing
the standard Bardeen approximation [equations
2and3in(23)] and the calculated partial density
of states (DOS) for CO molecules on the cluster
and on the tip (fig. S6) (23). The calculation re-
veals that the CO molecules on the corners indeed
have greater tunneling contributions than the
CO molecules on the edges, qualitatively explain-
ing the experimentally observed contrast of the
STMimageswiththesixbrightspotsplusone
in the center, as shown in Fig. 2, B and D. We
could explain the threefold symmetry of some of
the 19 atom clusters by adding three Cu atoms at
the center of the 19 atoms. These three low-
coordinated Cu atoms, producing the bright
center of the cluster images, can bind three ad-
ditional CO molecules and distort the tilt angles
of the peripheral CO molecules, as shown in Fig.
2, C and E [(details are shown in (23)].
Finally, we investigated the effect of clustering
on surface reactivity for the WGS reaction (i.e.,
CO+H
2
OCO
2
+H
2
), which Cu catalyzes. Water
does not adsorb on the Cu(111) surface at room
temperature (Fig. 4B) (32), whereas it dissocia-
tivelyadsorbsonthemoreactiveCu(110)surface
(32).OncethegasphaseCOat1Torrwaspumped
away, the STM images revealed that the Cu
clusters were still present, although atomic res-
olution could not be achieved, likely because
of the absence of CO molecules adsorbed on the
tip in high vacuum (Fig. 3A). In the presence of
2×10
9
Torr of H
2
O, the cluster-covered sur-
face was very active in dissociating water, as
shown by the increasing oxygen peak in both
the Auger electron spectra (AES) shown in Fig.
3B, and in the XPS spectra shown in Fig. 3C.
The APXPS spectrum indicates that the O peak
is a result of the dissociative adsorption of H
2
O
(Fig. 4A) and that no such peak appears after
experiments at 0.1 Torr of CO because clustering
of the Cu did not occur at lower CO pressures
(Figs. 1B and 3B). A similar effect was also ob-
served during exposure to CO+H
2
Omixtures,as
shown in Fig. 4A. The pristine Cu(111) surface, on
the other hand, not pre-exposed to CO, is inactive
(Fig. 4B).
Our findings open the possibility that other
soft materials (e.g., Ag, Au, and Zn) can similarly
undergo large reconstructions at sufficiently high
pressures of CO (or other molecules). We have also
demonstrated that the inactive (111) face of Cu for
water dissociation, a key step in the water-gas shif t
reaction, becomes highly activated as a result of
the CO-induced clustering. The need for this type
of study to extend our understanding of the working
ofcatalystsunderoperatingconditionsisclear.
REFERENCES AND NOTES
1. G. A. Somorjai, Introduction to Surface Chemistry and Catalysis
(Wiley-VCH, 1999).
2. G. Ertl, Angew. Chem. Int. Ed. 47, 35243535 (2008).
3. B. J. McIntyre, M. Salmeron, G. A. Somorjai, Rev. Sci. Instrum.
64, 687691 (1993).
4. L. Österlund et al., Phys. Rev. Lett. 86, 460463 (2001).
5. E. Laegsgaard et al., Rev. Sci. Instrum. 72, 35373542
(2001).
6. F. Tao, D. Tang, M. Salmeron, G. A. Somorjai, Rev. Sci. Instrum.
79, 084101 (2008).
7. F. Besenbacher, P. Thostrup, M. Salmeron, MRS Bull. 37,
677681 (2012).
8. C. T. Herbschleb et al., Rev. Sci. Instrum. 85, 083703
(2014).
9. M. Salmeron, R. Schlögl, Surf. Sci. Rep. 63, 169199
(2008).
10. M. Salmeron, MRS Bull. 38, 650657 (2013).
11. K. Klier, in Advances in Catalysis, Vol. 31,D.D.Eley,H.Pines,P.B.Weisz,
Eds. (Academic Press, 1982), pp. 243313.
12. D. S. Newsome, Catal. Rev. 21, 275318 (1980).
13. J. Szanyi, D. W. Goodman, Catal. Lett. 10, 383390 (1991).
14. J. Yoshihara, C. T. Campbell, J. Catal. 161, 776782
(1996).
15. M. Behrens et al., Science 336, 893897 (2012).
16. G. A. Olah, Angew. Chem. Int. Ed. 52, 104107 (2013).
17. M. Behrens, Angew. Chem. Int. Ed. 53, 1202212024 (2014).
18. K. Kambe, Phys. Rev. 99, 419422 (1955).
19. C. Kittel, Introduction to Solid State Physics, 8th Edition
(Wiley, 2005).
20. S. R. Longwitz et al., J. Phys. Chem. B 108, 144971450 2
(2004).
21. D. Tang, K. S. Hwang, M. Salmeron, G. A. Somorjai, J. Phys.
Chem. B 108, 1330013306 (2004).
22. F. Tao et al., Science 327, 850853 (2010).
23. See supplementary materials on Science Online
24. B. Eren et al., J. Phys. Chem. C 119, 1466914674 (2015).
25. J. Lagoute, X. Liu, S. Fölsch, Phys. Rev. Lett. 95,136801
(2005).
26. O. V. Lysenko, V. S. Stepanyuk, W. Hergert, J. Kirschner, Phys.
Rev. Lett. 89, 126102 (2002).
27. M. Mehlhorn, H. Gawronski, K. Morgenstern, Phys. Rev. Lett.
104, 076101 (2010).
28. H. J. Yang, T. Minato, M. Kawai, Y. Kim, J. Phys. Chem. C 117,
1642916437 (2013).
29. L. Bartels, D. Meyer, K. H. Rieder, Surf. Sci. Lett. 432,
L621L626 (1999).
30. M. Poensgen, J. F. Wolf, J. Frohn, M. Giesen, H. Ibach, Surf. Sci.
274, 430440 (1992).
31. M. M. Waldrop, Science 234, 673674 (1986).
32. S. Yamamoto et al., J. Phys. Chem. C 111, 78487850
(2007).
ACKNO WLED GME NTS
This work was supported by the Office of Basic Energy Sciences
(BES), Division of Materials Sciences and Engineering, of the U.S.
Department of Energy (DOE) under contract no. DE-AC02-
05CH11231, through the Chemical and Mechanical Properties of
Surfaces, Interfaces and Nanostructures program (FWP KC3101).
It used resources of the National Energy Research Scientific
Computing Center and the Advanced Light Source, which are
supported by the Office of Science of the U.S. DOE. The
computation used resources from the Oak Ridge Leadership
Computing Facility (OLCF), with time allocated by the Innovative
and Novel Computational Impact on Theory and Experiment
(INCITE) project.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/351/6272/475/suppl/DC1
Materials and Methods
Supplementary Text
Figs. S1 to S8
Tables S1 to S3
References (3341)
16 November 2015; accepted 14 December 2015
10.1126/science.aad8868
OCEANOGRAPHY
Enhanced East Pacific Rise
hydrothermal activity during the last
two glacial terminations
D. C. Lund,
1
*P. D. Asimow,
2
K. A. Farley,
2
T. O. Rooney,
3
E. Seeley,
1
E. W. Jackson,
4
Z. M. Durham
4
Mid-ocean ridge magmatism is driven by seafloor spreading and decompression melting
of the upper mantle. Melt production is apparently modulated by glacial-interglacial
changes in sea level, raising the possibility that magmatic flux acts as a negative feedback
on ice-sheet size. The timing of melt variability is poorly constrained, however, precluding
a clear link between ridge magmatism and Pleistocene climate transitions. Here we present
well-dated sedimentary records from the East Pacific Rise that show evidence of enhanced
hydrothermal activity during the last two glacial terminations.We suggest that glacial maxima
and lowering of sea level caused anomalous melting in the upper mantle and that the
subsequent magmatic anomalies promoted deglaciation through the release of mantle heat
and carbon at mid-ocean ridges.
Sea leveldriven pressure variations due to
the growth and decay of ice sheets likely
modulate melt production in the upper
mantle on Milankovitch time scales (1,2).
Model simulations suggest that the mag-
nitude of the resulting signal at mid-ocean ridges
depends on the plate spreading rate, the melt
extraction velocity, and the thermal properties
of the lithosphere (1,3). Because of the slow
rate of melt migration in the upper mantle, the
magmatic signal at ridges probably lags changes
in sea level by thousands of years (1). Surveys of
ridge bathymetry reveal Milankovitch-scale fre-
quencies in abyssal-hill spacing, consistent with
the sea-level hypothesis (3,4). Bathymetry records
are subject to geological damping effects and
478 29 JANUARY 2016 VOL 351 ISSUE 6272 sciencemag.org SCIENCE
1
Deptartment of Marine Sciences, University of Connecticut,
Groton, CT 06340, USA.
2
Division of Geological and
Planetary Sciences, California Institute of Technology,
Pasadena, CA 91125, USA.
3
Department of Geological
Sciences, Michigan State University, East Lansing, MI 48824,
USA.
4
Department of Earth and Environmental Sciences,
University of Michigan, Ann Arbor, MI 48109, USA.
*Corresponding author. E-mail: david.lund@uconn.edu
RESEARCH |REPORTS
on February 2, 2016Downloaded from on February 2, 2016Downloaded from on February 2, 2016Downloaded from on February 2, 2016Downloaded from on February 2, 2016Downloaded from
substantial age uncertainties, however, and they
therefore require validation with other proxies.
Because hydrothermal activity along ridge sec-
tions is ultimately driven by magmatic heat, sed-
imentary records of hydrothermal output can
be used to assess the sea-level hypothesis and
determine the timing of magmatic anomalies
relative to key Pleistocene climate transitions.
The southern East Pacific Rise (SEPR) has the
fastest spreading rate and the highest magmatic
budget of any ridge in the global mid-ocean ridge
system (5). Due to its elevated magmatism, the
SEPR has over 50 known active vent sites from
5°S to 37°S (6), consistent with the global trend
in plume incidence versus magmatic budget for
ridges spanning a range of spreading rates (5,7).
Intense hydrothermal venting and topographical-
ly steered flow of plumes along the SEPR create
a spatially integrated pattern of metalliferous sed-
iments near the ridge crest (810). Compared
with slower ridges, SEPR sediments have anoma-
lously high metal concentrations (8,11), suggesting
that magmatism is the primary factor governing
hydrothermal input to these sedimentary archives
on geologic time scales. Hydrothermal plume par-
ticles are highly enriched in elements that are
derived directly from vents and scavenged from
seawater. Variations in the flux of these elements
to ridge-flank sediments should therefore reflect
long-term changes in hydrothermal activity.
We used a multiproxy geochemical strategy to
reconstruct SEPR hydrothermal activity during
the last glacial cycle. We analyzed a total of sev-
en ridge-crest cores from 6°S and 11°S, whe re
the h alf -spreading rate ave rages 75 mm/year
(Fig. 1). Together with two published records from
near the East Pacific Rise (EPR)Dietz volcanic-
ridge triple junction at 1°N (12), the locations
span a range of spreading rates, sedimentary en-
vironments, and surface-ocean productivity re-
gimes. To control for spatial heterogeneity in
plume incidence versus magmatic budget (7), the
sampling locations span three separate EPR seg-
ments. At each segment, we analyzed cores from
both sides of the ridge axis to address potential
biases due to horizontal sediment focusing, bio-
turbation, and spatial variability in hydro thermal-
plume direction. Radiocarbon and oxygen isotopic
analyses of planktonic foraminifera provided age
control for each core (13). Major and trace ele-
ment concentrations were determined using x-ray
fluorescence (XRF) and inductively coupled plas-
ma mass spectrometry (ICP-MS). The fluxes of
hydrothermal components were estimated using
both mass accumulation rates and the
3
He nor-
malization method (13). Given that plume parti-
cles primarily consist of Fe oxyhydroxides and
Mn oxides (14), we used sedimentary Fe and Mn
to track hydrothermal inputs. To cross-check the
Fe and Mn results, we also measured arsenic, whi ch
is scavenged from seawater by Fe oxyhydroxides
and varies coherently with Fe in hydrothermal-
plume particles (15)andSEPRsediments(16).
Oxygen stable isotope records from 1°N, 6°S,
and 11°S outline marine isotope stages 1, 2, and 3,
indicating that there has been minimal strati-
graphic disturbance of the cores due to sediment
winnowing or downslope transport (Fig. 2). The
flux of Fe in all nine records peaks between
10 and 20 thousand years before the present
(ky B.P.). Manganese fluxes to EPR sediments
follow a similar pattern, with maximum values
centered at ~15 ky B.P. Arsenic fluxes at 6°S
and 11°S reach a maximum between 10 and 20 ky
B.P., supporting the Fe and Mn results. Offsets
between time series are generally 5 ky or less
(Fig. 2), similar to the age uncertainty associa-
ted with the mass accumulation rate method (13).
Results from the
3
He normalization technique,
which yields fluxes that are insensitive to age-
model uncertainty, show that positive shifts in
metal fluxes at 11°S and 6°S occurred within 2 ky
of one another (fig. S1). Thus, the overall pattern
for the past 50 ky is one of coherent variations
in hydrothermal sedimentation along 1300 km
of the EPR, with maximum metal inputs coincid-
ing with the last deglaciation (Termination I).
Two cores at 11°S span the penultimate de-
glaciation (Termination II), including core Y71-
07-53 on the western flank of the SEPR and core
Y71-07-47 on the eastern flank (Fig. 3). Metal
fluxes are higher in the western-flank core, con-
sistent with the east-west contrast in the shorter
records (Fig. 2) and the spatial pattern in metal
concentrations of late Holocene sediments (8).
In the western-flank core, the flux of each metal
increases markedly at ~140 ky B.P., reaches a
maximum by 130 ky B.P., and then returns to
backgroundlevelsby120kyB.P.(Fig.3).Asim-
ilarpatternoccursintheeastern-flankcore.The
contemporaneous signal at these loca tions indi-
cates that hydrothermal inputs on each side of
the ridge crest varied in phase. The records also
show that the maximum flux of hydrothermal
metals coincided with Termination II, similar to
the pattern for Termination I.
Diagenetic overprinting, horizontal sediment
focusing, and dilution with nonhydrothermal
components can complicate the interpretation
of hydrothermal proxies. Diagenetic remobili-
zation should influence Fe oxyhydroxides and
MnO
2
differently, given the large offset in their
redox potentials (17), yet we observed coherent
down-corevariationsinFe,Mn,andAs,regard-
less of location. Furthermore, down-core Fe/Mn
ratios generally fall within the expected range
for hydrothermal input (8). Anomalously high
SCIENCE sciencemag.org 29 JANUARY 2016 VOL 351 ISSUE 6272 479
x
+
KLH068
KLH093
68 mm/yr
17 mm/yr
49 mm/yr
79 mm/yr
75 mm/yr
Pacific
Plate
Nazca
Plate
Y71-07-49
Y71-07-51
11.0°S
11.5°S
107°W108°W
6.0°S
6.5°S
Y71-09-115
Y71-09-106
Y71-09-104
Y71-07-53
Y71-07-47
Bathymetry
location
110°W111°W
1°N Sites
6°S Sites
11°S Sites
Fig. 1. Map of sampling sites near the EPR. Core locations include 1°N (cores KLH068 and KLH093)
(12), 6°S (core Y71-09-104 in blue, Y71-09-106 in red, and Y71-09-115 in black), and 11°S (cores Y71-07-47
and Y71-07-49 in magenta and Y71-07-51 and Y71-07-53 in black). Also shown is the location of the
bathymetry record at 17°S (4). Half-spreading rates are shown in white (www.ldeo.columbia.edu/~menke/
plates.html). The half-spreading rate at 1°N (17 mm/year measured at Dietz volcanic ridge) is from (29).
The map was generated using GeoMapApp (www.geomapapp.org).
RESEARCH |REPORTS
Fe/Mn ratios at 1°N are probably due to
suboxic diagenesis and Mn remobilization
(13). At the 1°N locations, near-zero Mn levels
before 20 ky BP are likely driven by MnO
2
reduction and upward migration of dissolved
Mn
2+
. Nevertheless, the overall coherent pat-
tern in Fe records from the high-productivity
equatorial Pacific (1°N) to the northern edge of
the subtropical gyre (11°S) indicates that the
organic carbon flux to the sediments is not a
first-order control on down-core metal varia-
bility. Sediment focusing is an equally unlikely
explanation, given the similar pattern in
multiple cores from a range of sedimentary
environments. Focusing factors estimated using
3
He also show no evidence for anomalous
horizontal sediment transport during Termi-
nation I (fig. S3). Lastly, the
3
He-based metal
fluxes are consistent with the mass accu-
mulation rate results, indicating that carbon-
ate dilution was not a primary driver of the
down-core signal. Taken together, these lines
of evidence indicate that the metal fluxes
primarily reflect the input of hydrothermal-
plume particles to ridge-crest sediments.
The temporal variability in metal fluxes is sim-
ilar to that in seafloor bathymetry at 17°S on
the SEPR, implying that both have a common
driver (Fig. 4). A lowering of sea level due to
ice-sheet expansion would promote decompres-
sion melting in the upper mantle. The resulting
increase in melt delivery to the ridge crest should
result in shoaling of the bathymetry and greater
hydrothermal activity (1,3). Ice-sheet retreat and
rising sea level would have the opposite effect.
Shallower bathymetry on the SEPR generally
corresponds to elevated hydrothermal fluxes,
consistent with the expected pattern (Fig. 4).
The bathymetry record lags the hydrothermal
proxies by ~10 ky, however (fig. S4). The offset is
most likely due to age uncertainty in the bathy-
metrytimeseries,wheretheagemodelisbased
on a half-spreading rate that optimizes the match
between the bathymetry and atmospheric CO
2
records (4). More generally, age constraints for
late Pleistocene oceanic crust are limited to two
control points, an assumed zero age at the ridge
crest and the Brunhes-Matuyama boundary at
780 ky B.P. Even if reliable absolute ages were
available for individual abyssal hills, it is unlikely
that their bathymetry would reflect only the melt
delivery that occurred when that oceanic crust
was at the ridge crest, because of the confound-
ing influences of lower crustal accretion, surface
480 29 JANUARY 2016 VOL 351 ISSUE 6272 sciencemag.org SCIENCE
0
20
40
60
0
5
10
15
0 10 20 30 40 50
0
0.01
0.02
20
30
40
50
50
100
150
5
10
15
10
20
30
40
0 10 20 30 40 50
0
0.02
0.04
0
0.05
0.1
0.15
0 10 20 30 40 50
0
5
10
15
KLH068
KLH093
δ18O (‰)
Mn flux (µg cm-2 yr-1)
−1.5
−1
−0.5
0
0.5
1
123
Calendar Age (kyr BP) Calendar Age (kyr BP)
Calendar Age (kyr BP)
Y71-09-106
Y71-09-104
Y71-09-115
123
Y71-09-115 * 0.5
Y71-07-49
Y71-07-51
12 3
EPR 6°S EPR 11°S
EPR 1°N
As flux (mg cm-2 yr-1) Fe flux (µg cm-2 yr-1)Mn flux (µg cm-2 yr-1)δ18O (‰)
−2
−1.5
−1
−0.5
0
0.5
−2
−1.5
−1
−0.5
0
0.5
As flux (µg cm-2 yr-1) Fe flux (µg cm-2 yr-1)Mn flux (µg cm-2 yr-1)δ18O (‰)
10
20
30
40
50
Fe flux (µg cm-2 yr-1)
Fig. 2. Metal fluxes for the past 50 ky at three EPR
segments located at 1°N, 6°S, and 11°S. Flux esti-
mates are based on mass accumulation rates and metal
concentrations (13). Data from 1°N (12)areshownin
the first column, including (A) planktonic d
18
O, (B)Fe
flux, and (C) Mn flux. Data from 6°S are shown in the
second column, including (D) planktonic d
18
O, (E)Fe
flux, (F)Mnflux,and(G) As flux. Data from 11°S are
shown in the third column, including (H) planktonic
d
18
O, (I)Feflux,(J)Mnflux,and(K)Asflux.Maximum
fluxes occur during Termination I (gray vertical bar).
Typical errors for each record (crosses in each panel) represent the uncertainty of the flux estimates (vertical line) and age model error (horizontal line).
Calendar-corrected radiocarbon ages are shown as triangles. Approximate time intervals for marine isotope stages 1 to 3 are indicated in the top row of panels.
Arsenic results are not available for cores collected at 1°N because these cores were analyzed using XRF rather than ICP-MS (12).
RESEARCH |REPORTS
lava flows, and vertical and horizontal offsets of
crustal blocks by faulting (3,4,18,19). Although
bathymetric time series are useful for identifying
Milankovitch frequencies, the absolute timing
of events is poorly constrained by these records.
Hydrothermal proxies, on the other hand, can
be accurately dated using radiocarbon and oxy-
gen isotope stratigraphy. As a result, we are able
to infer that intervals of intense hydrothermal
activity on the EPR occurred during the last two
glacial terminations.
The coincidence in timing between hydrother-
mal maxima and glacial terminations implies
that there may be a direct causal relationship
between sea-level rise and hydrothermal activity.
Our understanding of the physical mechanisms
of decompression melting and melt migration
to the ridge axis suggests a more complex rela-
tionship, however. Proxies of magmatic flux should
lag sea-level changes by thousands of years, be-
cause of the slow rate of melt migration from the
magma source region to the ridge axis (1). During
the Last Glacial Maximum, the maximum rate of
sea-level decrease (and hence of pressure release
in the melting regime) occurred between 30 and
25 ky B.P. (20), or 15 ± 5 ky before the inferred
maximum in EPR hydrothermal activity (Fig. 2).
We observed a similar lag between the maximum
rate of sea-level rise at ~15 ky B.P. (20)andthe
late Holocene minimum in metal flux. Assuming
an average melt origin depth of 50 km (21), the
implied melt extraction velocities range from
2.5 to 5 m/year, which is consistent with the rate
of >1 m/year implied by U/Th disequilibrium
in zero-age mid-ocean ridge basalts (22) but
much lower than the estimates of >50 m/year
based on the time lag between deglaciation and
volcanism in Iceland (23).Ourestimateisinde-
pendent of previous methods and provides a
range of constraints for refining models of melt
extraction at fast spreading centers.
Our results support the hypothesis that en-
hanced ridge magmatism, hydrothermal output,
and perhaps mantle CO
2
flux act as a negative
feedback on ice-sheet size (1,4). Although the
modern carbon output from ridges is small (0.02
to 0.2 Pg C/year) (24), the flux probably increased
as a result ofsea-level modulation. Carbon sources
at off-axis locations, backarc basins, and island
arcsmayalsoamplifythemid-oceanridgesignal
(2). The long melt-migration times for carbon-
rich melts may lead to considerable differences
in timing between hydrothermal and carbon-flux
variations, however (25). Another mechanism
whereby magmatic variations may influence
climate is the hydrothermal heat flux itself. En-
hanced geothermal heat flux should warm and
destabilize the deep ocean (26), with excess heat
emerging along isopycnals into the surface South-
ern Ocean (26,27). Temperatures in the deep
SCIENCE sciencemag.org 29 JANUARY 2016 VOL 351 ISSUE 6272 481
Fig. 3. Planktonic d
18
O
and metal fluxes
spanning Termination
II at 11°S. The time
series are from the
eastern (core Y71-07-47;
magenta) and western
(core Y71-07-53; black)
flanks of the EPR.
(A) Planktonic d
18
O
results superimposed
on a global benthic
d
18
Ostack[LR04(30)]
(gray line), (B)Feflux,
(C)Mnflux,and(D)As
flux. In (A), thin lines
indicate data from dis-
crete samples, and thick
lines indicate time series
smoothed with a three-
point running mean. In
(B) to (D), flux estimates
are based on mass
accumulation rates and
metal concentrations.
Glacial terminations are
indicatedbydashedver-
tical lines. Hydrothermal
metal fluxes peaked dur-
ing Termination II (T2).
Error bars for the west-
ern flank (gray shaded
area) and eastern flank
(thin magenta lines)
reflect the uncertainty of
the flux estimates in (B)
to (D) (13). Metal data
for core Y71-07-53 (16)
were assigned ages
based on the d
18
Ostra-
tigraphy generated for
this work. Arsenic data
are not available for core
Y71-07-47 because it
was analyzed using XRF
rather than ICP-MS.
3
3.5
4
4.5
5
5.5
−1.5
−1
−0.5
0
0.5
40
80
120
160
10
20
30
40
0 50 100 150 200
0
0.1
0.2
0.3
0.4
Calendar Age (kyr BP)
LR04 δ18O Stack (‰)
Planktonic δ18O (‰)
Fe flux (µg cm−2 yr−1)Mn flux (µg cm−2 yr−1)As flux (µg cm−2 yr−1)
Mn flux (µg cm−2 yr−1)Fe flux (µg cm−2 yr−1)
0
10
20
30
40
50
2
4
6
8
10
12
14
0
western
flank
eastern
flank
T2T1
0 50 100 150 200
−2
−1
0
1
2
3
4
Calendar Age (kyr BP)
Normalized Iron
−80
−60
−40
−20
0
20
40
60
0 50 100 150 200
−2
−1
0
1
2
3
4
Calendar Age (kyr BP)
Normalized Manganese
−80
−60
−40
−20
0
20
40
60
0 50 100 150 200
−2
−1
0
1
2
3
4
Calendar A
g
e
(
k
y
r BP
)
Normalized Arsenic
−80
−60
−40
−20
0
20
40
60
17°S Bathymetry Anomaly (m) 17°S Bathymetry Anomaly (m) 17°S Bathymetry Anomaly (m)
Fig. 4. Normalized metal fluxes at 11°S com-
paredwithEPRbathymetry.The hydrothermal
time series are from the eastern (magenta) and
western (black) flanks of the EPR and include (A)
Fe fl ux, ( B)Mnflux,and(C) As flux. We normalized
each record by subtracting the mean and dividing
by the standard deviation of each time series to
facilitate comparison between cores with differ-
ent mean metal concentrations. The results include
both discrete samples (thin lines) and time series
smoothed with a 20-ky-wide Gaussian window
(thick lines) to approximate the resolution of the
bathymetry compilation at 17°S (gray lines) (4).
Fluxes from 0 to 40 ky are based on the results
from Fig. 2; the interval from 40 to 200 ky B.P. is
based on results shown in Fig. 3.
RESEARCH |REPORTS
eastern tropical Pacific and Antarctica peaked
during each of the last two glacial terminations
(28), consistent with the timing of enhanced EPR
hydrothermal activity.
Isolating a mechanistic linkage between ridge
magmatism and glacial terminations will require
a suite of detailed proxy records from multiple
ridges that are sensitive to mantle carbon and
geothermal inputs, as well as modeling studies
of their influence in the ocean interior. The
EPR results establish the timing of hydrothermal
anomalies, an essential prerequisite for deter-
mining whether ridge magmatism can act as a
negative feedback on ice-sheet size. The data
presented here demonstrate that EPR hydro-
thermal output increased after the two largest
glacial maxima of the past 200,000 years, im-
plicating mid-ocean ridge magmatism in glacial
terminations.
REFERENCES AND NOTES
1. D. C. Lund, P. D. Asimow, Geochem. Geophys. Geosyst. 12,
Q12009 (2011).
2. P. Huybers, C. Langmuir, Earth Planet. Sci. Lett. 286, 479491
(2009).
3. J. W. Crowley, R. F. Katz, P. Huybers, C. H. Langmuir,
S. H. Park, Science 347, 12371240 (2015).
4. M. Tolstoy, Geophys. Res. Lett. 42, 13461351 (2015).
5. E. T. Baker, G. R. German, in Mid-Ocean Ridges: Hydrothermal
Interactions Between the Lithosphere and Oceans,
C. R. German, J. Lin, L. M. Parson, Eds. (Geophysical
Monograph Series vol. 148, American Geophysical Union,
2004), pp. 245266.
6. S. E. Beaulieu, E. T. Baker, C. R. German, A. Maffei, Geochem.
Geophys. Geosyst. 14, 48924905 (2013).
7. E. T. Baker, Geochem. Geophys. Geosyst. 10, Q06009 (2009).
8. J. Dymond, Geol. Soc. Am. 154, 133174 (1981).
9. G. B. Shimmield, N. B. Price, Geochim. Cosmochim. Acta 52,
669677 (1988).
10. K. G. Speer, M. E. Maltrud, A. M. Thurnherr, in Energy and Mass
Transfer in Hydrothermal Systems, P. E. Halbach, V. Tunnicliffe,
J. R. Hein, Eds. (Dahlem University Press, 2003), pp. 287302.
11. K. Boström, M. N. Peterson, O. Joensuu, D. E. Fisher, J. Geophys.
Res. 74,32613270 (1969).
12. M. F rank et al., Paleoceanography 9,559578
(1994).
13. Material and methods are available as supplemental materials
on Science Online.
14. C. R. German, S. Colley, M. R. Palmer, A. Khripounoff,
G. P. Klinkhammer, Deep Sea Res. Part I Oceanogr. Res. Pap.
49, 19211940 (2002).
15. R. R. Cave, C. R. German, J. Thomson, R. W. Nesbitt, Geochim.
Cosmochim. Acta 66, 19051923 (2002).
16. T. Schaller, J. Morford, S. R. Emerson, R. A. Feely, Geochim.
Cosmochim. Acta 64, 22432254 (2000).
17. S. Emerson, J. I. Hedges, in Treatise on Geochemistry,
K. K. Turekian, H. D. Holland, Eds. (Elsevier, vol. 6, 2004),
pp. 293319.
18. J. A. Goff, Science 349, 1065 (2015).
19. J. A. Olive et al., Science 350, 310313 (2015).
20. P. U. Clark et al., Science 325, 710714 (2009).
21. K. Key, S. Constable, L. Liu, A. Pommier, Nature 495, 499502
(2013).
22. P. B. Kelemen, G. Hirth, N. Shimizu, M. Spiegelman, H. J. B. Dick,
Philos. Trans. R. Soc. Lond. A 355,283318 (1997).
23. J. Maclennan, M. Jull, D. McKenzie, L. Slater, K. Gronvold,
Geochem. Geophys. Geosyst. 3,125 (2002).
24. P. Cartigny, F. Pineau, C. Aubaud, M. Javoy, Earth Planet.
Sci. Lett. 265, 672685 (2008).
25. J. M. A. Burley, R. F. Katz, Earth Planet. Sci. Lett. 426, 246258
(2015).
26. M. Hofmann, M. A. Morales Maqueda, Geophys. Res. Lett. 36,
L03603 (2009).
27. J. Emile-Geay, G. Madec, Ocean Sci. 5, 203217 (2009).
28. P. Martin, D. Archer, D. W. Lea, Paleoceanography 20, PA2015
(2005).
29. D. K. Smith, H. Schouten, L. Montési, W. Zhu, Earth Planet.
Sci. Lett. 371372,615 (2013).
30. L. E. Lisiecki, M. E. Raymo, Paleoceanography 20, PA1003
(2005).
ACKNOW LEDGM ENTS
We dedicate this paper to J. Dymond, whose 1981 treatise on Nazca
plate sediments made this work possible. We are also indebted
to the Oregon State University Core Repository for carefully
preserving the EPR sediment cores since they were collected in
the early 1970s. We are grateful to L. Wingate at the University of
Michigan and M. Cote at the University of Connecticut for
technical support. This work has benefited from discussions
with J. Granger, P. Vlahos, B. Fitzgerald, and M. Lyle. Data
presented here are available on the National Oceanic and
Atmospheric Administrations Paleoclimatology Data website
(www.ncdc.noaa.gov/data-access/paleoclimatology-data). Funding
was provided by the University of Michigan and the University
of Connecticut.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/351/6272/478/suppl/DC1
Materials and Methods
Supplementary Text
Figs. S1 to S11
Tables S1 to S5
References (3145)
14 September 2015; accepted 6 January 2016
10.1126/science.aad4296
HISTORY OF SCIENCE
Ancient Babylonian astronomers
calculated Jupiters position from the
area under a time-velocity graph
Mathieu Ossendrijver*
Theideaofcomputingabodys displacement as an area in time-velocity space is usually traced
back to 14th-century Europe. I show that in four ancient Babylonian cuneiform tablets, Jupiters
displacement along the ecliptic is computed as the area of a trapezoidal figure obtained by
drawing its daily displacement against time.This interpretation is prompted by a newly
discovered tablet on which the same computation is presented in an equivalent arithmetical
formulation. The tablets date from 350 to 50 BCE. The trapezoid procedures offer the first
evidence for the use of geometrical methods in Babylonian mathematical astronomy, which was
thus far viewed as operating exclusively with arithmetical concepts.
The so-called trapezoid procedures examined
in this paper have long puzzled historians
of Babylonian astronomy. They belong to
the corpus of Babylonian mathematical as-
tronomy, which comprises about 450 tab-
lets from Babylon and Uruk dating between 400
and 50 BCE. Approximately 340 of these tablets
are tables with computed planetary or lunar data
arranged in rows and columns (1). The remaining
110 tablets are procedure texts with computa-
tional instructions (2), mostly aimed at comput-
ing or verifying the tables. In all of these texts the
zodiac, invented in Babylonia near the end of the
fifth century BCE (3), is used as a coordinate sys-
tem for computing celestial positions. The un-
derlying algorithms are structured as branching
chains of arithmetical operations (additions, sub-
tractions, and multiplications) that can be rep-
resented as flow charts (2). Geometrical concepts
are conspicuously absent from these texts, whereas
they are very common in the Babylonian mathe-
matical corpus (47). Currently four tablets, most
likely written in Babylon between 350 and 50 BCE,
are known to preserve portions of a trapezoid
procedure (8). Of the four procedures, here labeled
B to E (figs. S1 to S4), one (B) preserves a men-
tion of Jupiter and three (B, C, E) are embedded
in compendia of procedures dealing exclusively
with Jupiter. The previously unpublished text D
probably belongs to a similar compendium for
Jupiter. In spite of these indications of a connec-
tion with Jupiter, their astronomical significance
was previously not acknowledged or understood
(1,2,6).
A recently discovered tablet containing an un-
published procedure text, here labeled text A (Fig. 1),
sheds new light on the trapezoid procedures. Text A
most likely originates from the same period and
location (Babylon) as texts B to E (8). It contains
a nearly complete set of instructions for Jupiters
motion along the ecliptic in accordance with the
so-called scheme X.S
1
(2). Before the discovery of
text A, this scheme was too fragmentarily known
for identifying its connection with the trapezoid
procedures. Covering one complete synodic cycle,
scheme X.S
1
begins with Jupiters heliacal rising
(first visible rising at dawn), continuing with its
first station (beginning of apparent retrograde
motion), acronychal rising (last visible rising at
dusk), second station (end of retrograde motion),
and heliacal setting (last visible setting at dusk)
(2). Scheme X.S
1
and the four trapezoid procedures
are here shown to contain or imply mathematically
equivalent descriptions of Jupitersmotionduring
the first 60 days after its first appearance. Whereas
scheme X. S
1
employs a purely arithmetical ter-
minology, the trapezoid procedures operate with
geometrical entities.
482 29 JANUARY 2016 VOL 351 ISSUE 6272 sciencemag.org SCIENCE
Excellence Cluster TOPOIInstitute of Philosophy, Humboldt
University, Berlin, Germany.
*Corresponding author. E-mail: mathieu.ossendrijver@hu-berlin.de
RESEARCH |REPORTS
DOI: 10.1126/science.aad4296
, 478 (2016);351 Science et al.D. C. Lund
glacial terminations
Enhanced East Pacific Rise hydrothermal activity during the last two
This copy is for your personal, non-commercial use only.
clicking here.colleagues, clients, or customers by , you can order high-quality copies for yourIf you wish to distribute this article to others
here.following the guidelines can be obtained byPermission to republish or repurpose articles or portions of articles
): February 1, 2016 www.sciencemag.org (this information is current as of
The following resources related to this article are available online at
/content/351/6272/478.full.html
version of this article at: including high-resolution figures, can be found in the onlineUpdated information and services,
/content/suppl/2016/01/27/351.6272.478.DC1.html
can be found at: Supporting Online Material
/content/351/6272/478.full.html#ref-list-1
, 6 of which can be accessed free:cites 41 articlesThis article
/cgi/collection/oceans
Oceanography /cgi/collection/geochem_phys
Geochemistry, Geophysics subject collections:This article appears in the following
registered trademark of AAAS. is aScience2016 by the American Association for the Advancement of Science; all rights reserved. The title CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on February 2, 2016Downloaded from
... The growing evidence shows that hydrothermal activities on MORs have responded to the global glacial-interglacial sea level changes, such as sediment records from the East Pacific Rise (Lund et al., 2016), the Juan de Fuca Ridge (Costa et al., 2017), the Mid-Atlantic Ridge (Middleton et al., 2016), and the CR (Li et al., 2023). It has been proposed that declining sea levels during glacial periods may have reduced hydrostatic pressure on the melt centers of MORs and increased melt production, leading to enhanced hydrothermal activity (Lund et al., 2016;Costa et al., 2017). ...
... The growing evidence shows that hydrothermal activities on MORs have responded to the global glacial-interglacial sea level changes, such as sediment records from the East Pacific Rise (Lund et al., 2016), the Juan de Fuca Ridge (Costa et al., 2017), the Mid-Atlantic Ridge (Middleton et al., 2016), and the CR (Li et al., 2023). It has been proposed that declining sea levels during glacial periods may have reduced hydrostatic pressure on the melt centers of MORs and increased melt production, leading to enhanced hydrothermal activity (Lund et al., 2016;Costa et al., 2017). Some evidence suggests that rapid changes in sea level affect the hydrothermal output on MORs (Middleton et al., 2016). ...
... Thus, hydrothermal event H2 is likely to be driven by the increased magmatic production associated with sea-level fall during the LGM. Previous studies of the hydrothermal records have shown that the peak of hydrothermal activity associated with the LGM may lag the onset of the LGM by 5-10 ka due to the time required for the migration of melts generated in the upper mantle to the heat reservoir (Lund et al., 2016;Li et al., 2023). However, the hydrothermal record in the Tianshi field appears to be an exception to this rule, with the peak in Fe H lagging the onset of the LGM by only ~1 ka (Fig.10). ...
Article
A new hydrothermal field (Tianshi) was discovered on the rift valley wall through plume anomaly surveys and geological work conducted in 2012 and 2018 between 2°35′N and 2°43′N of the slow-spreading Carlsberg Ridge (CR). Here, the results of two expeditions conducted to detect and characterize the new hydrothermal field are reported. Mineralogical and geochemical data, as well as 14C ages of a sediment core collected near the field are presented to reveal the hydrothermal history. Results show that the Tianshi field is a basalt-hosted hydrothermal system. Geochemical data of the sediments collected near the field indicate a strong hydrothermal contribution, and hydrothermal Fe and Cu fluxes range from 30 to 155 mg/(cm2.ka) and 0.59 to 11.49 mg/(cm2.ka), respectively. Temporal variations in the fluxes of hydrothermal Fe indicate that there have been at least three amplified hydrothermal venting events (H1, H2, and H3) in the Tianshi field over the last 35.2 ka, in 28.6–35.2 ka BP, 22.0–27.6 ka BP, and 1.2–11.4 ka BP, respectively. Hydrothermal event H2 was driven by an increased magmatic production associated with sea level fall during the Last Glacial Maximum, while event H3 was promoted by tectonic activity associated with a rapid sea level rise. This study further verified the role of sea level change in modulating hydrothermal activity on mid-ocean ridges.
... Marine sedimentary fluxes are important for understanding oceanic elemental output and input rates, as well as spatio-temporal patterns in the deposition of windblown dust (Winckler et al., 2008), biogenic minerals (Anderson et al., 2009), and hydrothermal metals (Lund et al., 2016). Typically, sediment mass accumulation rates (MARs) are calculated by combining estimates of sediment dry bulk density and linear sedimentation rates between dated stratigraphic tie points (e.g., (Lyle and Dymond, 1976). ...
... Our measured 3 He/ 4 He ratios in the South Pacific were exceptionally highat or above the conventionally-used extraterrestrial endmember 3 He/ 4 He of 2.4 x 10 -4 ( Fig. 2). This is consistent with previous measurements of 3 He/ 4 He ratios in the South Pacific (Lund et al., 2016(Lund et al., , 2019, and likely reflects a combination of low mass accumulation rates and low terrigenous inputs (Fig. 2). We thus use 4 x 10 -4 as our extraterrestrial 3 He/ 4 He endmember. ...
... White stars show locations of monocores (MC) collected during the GEO-TRACES GP16 expedition, and pink circles show locations of box cores collected during the SO245 expedition. Cyan squares show the locations of 3 He measurements made byLund et al. (2016Lund et al. ( , 2019. ...
... Recent data revealed the link between the hydrothermal activity and global climate changes [6,31,33]. The sea level decreases during global glaciation periods leading to a decrease in hydrostatic pressure. ...
... This suggestion is confirmed by a correlation of marine isotope stages, which register the warming and cooling stages, and temporal intervals of hydrothermal activity, which were distinguished from the study of seafloor sediment cores near the hydrothermal sites. It was found that the maxima of hydrothermal activity within the EPR and MAR approximately correspond to the two last cooling stages [31,33]. ...
... Also, fluctuations in sea level linked to glacio-eustatic cycles have been hypothesized to influence hydrothermal activity at mid-ocean ridges over the past 50 kyr (Middleton et al., 2016). The deposition of hydrothermal iron and copper is observed to have increased during the Last Glacial Maximum, followed by a rapid decline during the sea level rise associated with deglaciation (Lund et al., 2016;Middleton et al., 2016). The relationship between sea-level changes and volcanic activity also extends beyond the context of mid-ocean ridges to continental volcanic systems. ...
Article
Full-text available
Cyclic fluctuations in the frequency and intensity of volcanic activity are recorded during periods of global climate change. Volcano-sedimentary successions (e.g., in near-coastal environments) may reveal the interplay of glacio-eustatic fluctuations, controlling erosional vs. aggradational processes, and the pattern of volcanic activity. However, the idea of a causal link between Earth’s climate and volcanism is still debated, also because many prior studies have focused on a single glacial cycle. The strongest evidence for a connection between orbitally driven climate variations and volcanism lies in the observed periodicity of volcanic activity on a time-scale of 10^3–10^4 years parallel to glacial-interglacial climate fluctuations. This has suggested that volcanism may be influenced indirectly by Earth’s orbital factors, through their effects on climate and the resulting changes in the distribution of continental ice and seawater masses. The hypothesis of a glacio-eustatic control specifically connects Milankovitch cycles—such as the 100,000-year eccentricity cycle, and the 41,000-year obliquity cycle—to the frequency and intensity patterns of volcanic eruptions, as a result of crustal stress changes driven by the redistribution of ice masses and sea level fluctuations. The alternative hypothesis suggests a direct gravitational effect on the crustal stress field driven by orbital oscillations in Earth’s inclination and rotation. This would result into periodic intensifications of volcanic activity and related greenhouse gas emission, thus in turn influencing the intensity of Milankovitch periodicities on a global scale. Here, we present an overview of the ongoing debate on the cause-and-effect relationships of Earth’s orbital factors, periodic climate changes and volcanism. On these grounds, we point out possible research perspectives.
... The removal process need not be fundamentally different from the modern but could have been larger (see previous section). The most recent interruption for which there are data began around 25kyBP and peaked during deglaciation (Bryan et al., 2010;Lund and Asimow, 2011;Lund et al., 2016;Marchitto et al., 2007;Middleton et al., 2016;Rafter et al., 2019;Ronge et al., 2016;Stott et al., 2019a;Stott et al., 2019b). ...
Article
Mixing of high-temperature hydrothermal vent fluids with seawater leads to the precipitation of a large fraction of the non-conservative element load and scavenging of elements from seawater onto the particles formed. This substantially modifies hydrothermal fluxes into the ocean relative to the flux across the seafloor at vents. However, the particles formed in the plume are rarely in equilibrium within the uppermost sediments, and scavenged elements may be loosely bound to particle surfaces, meaning that the net flux from the ocean can be readily modified by processes operating during early diagenesis. Using bulk geochemistry, and progressive leaching experiments, we characterize sediments from around the Endeavour segment of the Juan de Fuca ridge to investigate scavenging and early diagenetic processes. Changes in bulk sediment composition with distance from the ridge (e.g., decreases in Fe/Ti) support a decreased hydrothermal input off-axis. Bulk-sediment and leachate data are interpreted as indicating substantial scavenging of elements such as P, V, Cr and REEs as previously suggested, with different elements scavenged with different efficiencies in different parts of the plume. Correlations of these elements with Fe in the HCl leachates, suggests their uptake is associated with Fe-oxyhydroxides. Water column scavenging is overprinted by fluxes associated with benthic scavenging for V and REE. In contrast, early diagenesis leads to complete loss of scavenged P back to the ocean. Leaching the samples with NaOH and acetic acid provides evidence for recrystallization of reactive (biogenic and hydrothermal) silica and CaCO3 during early diagenesis in all locations. The fraction of Fe and Mn leached by 1 M HCl in 1 h versus 24 h, and the amount leached in 24 h, vary systematically; these data can be explained by progressive recrystallization of hydrothermally-derived Fe and Mn minerals during early diagenesis. These early diagenetic reactions release almost all hydrothermally-derived Mn back into the ocean, while Fe remains largely fixed in the sediment. Large-scale loss of Co, Zn, Mo, Ag and Cd also occurs during early diagenesis, with some evidence As, Pb and U are also lost. In deeper, more reducing, sediments there is uptake of As, Mo, Ag, Cd and U, presumably via diffusion through the shallow oxic sediments. Such diagenetic processes mean that the net fluxes of elements into and out of the ocean associated with hydrothermal systems differ from those estimated from studies of either vent fluids or hydrothermal plumes.
Article
Recent studies have suggested a link between ice age sea level fluctuations and variations in magma production and crustal faulting along mid-ocean ridges based on the detection of Milankovitch cycle frequencies in topography off several ridges. These fluctuations have also been connected to variability in hydrothermal metal fluxes near ridges. Ice age sea level calculations have shown that the sea level change across glacial cycles will be characterized by significant geographic variability, that is, departures from eustasy, due to the gravitational, deformation and rotational effects of the glacial isostatic adjustment (GIA) process. Using a state-of-the-art GIA simulation that incorporates 3-D variations in Earth viscoelastic structure, including plate boundaries, and updated constraints on the magnitude and geometry of ice mass fluctuations, we predict global sea level changes from Last Glacial Maximum (LGM, 26 ka) to present and from the Penultimate Glacial Maximum (PGM, 143 ka) to the Last Interglacial (LIG, 128 ka). We focus on the results along three ridges: the Mid-Atlantic Ridge, Juan de Fuca Ridge and East Pacific Rise, which are examples of slow, intermediate and fast spreading ridges, respectively. Sea level change across the Mid-Atlantic Ridge shows the greatest variability, ranging from a sea level fall greater than 200 m in Iceland to a maximum rise of ∼150 m in the South Atlantic, with significant non-monotonicity north of the Equator as the ridge weaves across the field of sea level changes. We also calculate changes in crustal normal stress from LGM to present-day across the Mid-Atlantic and Juan de Fuca Ridges and the East Pacific Rise. These results indicate that the contribution from ice mass changes to the crustal stress field can be significant well away from the location of ancient ice complexes. We conclude that any exploration of the hypothesized links to magma production and crustal faulting must consider both ocean and ice loading effects and, more generally, the profound geographic variability of the GIA process.
Article
Full-text available
Ocean dissolved oxygen (DO) can provide insights on how the marine carbon cycle affects global climate change. However, the net global DO change and the controlling mechanisms remain uncertain through the last deglaciation. Here, we present a globally integrated DO reconstruction using thallium isotopes, corroborating lower global DO during the Last Glacial Maximum [19 to 23 thousand years before the present (ka B.P.)] relative to the Holocene. During the deglaciation, we reveal reoxygenation in the Heinrich Stadial 1 (~14.7 to 18 ka B.P.) and the Younger Dryas (11.7 to 12.9 ka B.P.), with deoxygenation during the Bølling-Allerød (12.9 to 14.7 ka B.P.). The deglacial DO changes were decoupled from North Atlantic Deep Water formation rates and imply that Southern Ocean ventilation controlled ocean oxygen. The coherence between global DO and atmospheric CO 2 on millennial timescales highlights the Southern Ocean’s role in deglacial atmospheric CO 2 rise.
Article
Full-text available
The geological record shows links between glacial cycles and volcanic productivity, both subaerially and at mid-ocean ridges. Sea-level-driven pressure changes could also affect chemical properties of mid-ocean ridge volcanism. We consider how changing sea-level could alter the \cotwo{} emissions rate from mid-ocean ridges, on both the segment and global scale. We develop a simplified transport model for a highly incompatible element through a homogenous mantle; variations in the melt concentration the emission rate of the element are created by changes in the depth of first silicate melting. The model predicts an average global mid-ocean ridge \cotwo{} emissions-rate of 53~Mt/yr, in line with other estimates. We show that falling sea level would cause an increase in ridge \cotwo{} emissions with a lag of about 100~kyrs after the causative sea level change. The lag and amplitude of the response are sensitive to mantle permeability and plate spreading rate. For a reconstructed sea-level time series of the past million years, we predict variations of up to 12\% (7~Mt/yr) in global mid-ocean ridge \cotwo{} emissions. The magnitude and timing of the predicted variations in \cotwo{} emissions suggests a potential role for ridge carbon emissions in glacial cycles.
Article
Full-text available
Glacial cycles redistribute water between oceans and continents, causing pressure changes in the upper mantle, with consequences for the melting of Earth's interior. Using Plio-Pleistocene sea-level variations as a forcing function, theoretical models of mid-ocean ridge dynamics that include melt transport predict temporal variations in crustal thickness of hundreds of meters. New bathymetry from the Australian-Antarctic ridge shows statistically significant spectral energy near the Milankovitch periods of 23, 41, and 100 thousand years, which is consistent with model predictions. These results suggest that abyssal hills, one of the most common bathymetric features on Earth, record the magmatic response to changes in sea level. The models and data support a link between glacial cycles at the surface and mantle melting at depth, recorded in the bathymetric fabric of the sea floor. Copyright © 2015, American Association for the Advancement of Science.
Article
Crowley et al. (Reports, 13 March 2015, p. 1237) propose that abyssal hill topography can be generated by variations in volcanism at mid-ocean ridges modulated by Milankovitch cycle-driven changes in sea level. Published values for abyssal hill characteristic widths versus spreading rate do not generally support this hypothesis. I argue that abyssal hills are primarily fault-generated rather than volcanically generated features. Copyright © 2015, American Association for the Advancement of Science.
Article
Recent studies have proposed that the bathymetric fabric of the seafloor formed at mid-ocean ridges records rapid (23,000 to 100,000 years) fluctuations in ridge magma supply caused by sealevel changes that modulate melt production in the underlying mantle. Using quantitative models of faulting and magma emplacement, we demonstrate that, in fact, seafloor-shaping processes act as a low-pass filter on variations in magma supply, strongly damping fluctuations shorter than about 100,000 years. We show that the systematic decrease in dominant seafloor wavelengths with increasing spreading rate is best explained by a model of fault growth and abandonment under a steady magma input. This provides a robust framework for deciphering the footprint of mantle melting in the fabric of abyssal hills, the most common topographic feature on Earth.
Chapter
The “magmatic budget hypothesis” proposes that variability in magma supply is the primary control on the large-scale hydrothermal distribution pattern along oceanic spreading ridges. The concept is simple but several factors make testing the hypothesis complex: scant hydrothermal flux measurements, temporal lags between magmatic and hydrothermal processes, the role of permeability, nonmagmatic heat sources, and the uncertainties of vent-field exploration. Here we examine this hypothesis by summarizing our current state of knowledge of the global distribution of active vent fields, which presently number ~280, roughly a quarter of our predicted population of ~1000. Approximately 20% of the global ridge system has now been surveyed at least cursorily for active sites, but only half that length has been studied in sufficient detail for statistical treatment. Using 11 ridge sections totaling 6140 km we find a robust linear correlation between either site frequency or hydrothermal plume incidence and the magmatic budget estimated from crustal thickness. These trends cover spreading rates of 10-150 mm/yr and strongly support the magma budget hypothesis. A secondary control, permeability, may become increasingly important as spreading rates decrease and deep faults mine supplemental heat from direct cooling of the upper mantle, cooling gabbroic intrusions, and serpentinization of underlying ultramafics. Preliminary observations and theory suggest that hydrothermal activity on hotspot-affected ridges is relatively deficient, although paucity of data precludes generalizing this result. While the fullness of our conclusions depends upon further detailed study of vent field frequency, especially on slow-spreading ridges, they are consistent with global distributions of deep-ocean ³He, an unequivocally magmatic tracer.
Article
Seafloor eruption rates, and mantle melting fueling eruptions, may be influenced by sea-level and crustal loading cycles at scales from fortnightly to 100 kyr. Recent mid-ocean ridge eruptions occur primarily during neap tides and the first 6 months of the year, suggesting sensitivity to minor changes in tidal forcing and orbital eccentricity. An ~100 kyr periodicity in fast-spreading seafloor bathymetry, and relatively low present-day eruption rates, at a time of high sea-level and decreasing orbital eccentricity suggest a longer term sensitivity to sea-level and orbital variations associated with Milankovitch cycles. Seafloor spreading is considered a small but steady contributor of CO2 to climate cycles on the 100 kyr time scale, however this assumes a consistent short-term eruption rate. Pulsing of seafloor volcanic activity may feed back into climate cycles, possibly contributing to glacial/inter-glacial cycles, the abrupt end of ice ages, and dominance of the 100 kyr cycle.
Article
Preface. Introduction. 1 Surfaces: An Introduction. 2 The Structure of Surfaces. 3 Thermodynamics of Surfaces. 4 Dynamics at Surfaces. 5 Electrical Properties of Surfaces. 6 Surface Chemical Bond. 7 Mechanical Properties of Surfaces. 8 Polymer Surfaces and Biointerfaces. 9 Catalysis by Surfaces. Index.