Auxiliary text for

advertisement
1
Auxiliary text for
2
3
Geochemistry of volcanic glasses from the Louisville Seamount Trail (IODP
4
Expedition 330): implications for eruption environments and mantle melting
5
Alexander R.L. Nichols1#, Christoph Beier2, Philipp A. Brandl2, Stefan H. Krumm2
6
7
8
1
9
Science and Technology (JAMSTEC), 2-15 Natsushima-cho, Yokosuka, Kanagawa,
Institute for Research on Earth Evolution (IFREE), Japan Agency for Marine Earth
10
237-0061 Japan
11
2
12
Schlossgarten 5, 91054 Erlangen, Germany
GeoZentrum
Nordbayern,
13
14
15
email: nichols@jamstec.go.jp
Friedrich-Alexander-Universität
Erlangen-Nürnberg,
16
1. Drill Sites
17
1.1 Canopus Guyot (Site U1372):
18
Site U1372 is on the summit plain of the northern volcanic center of Canopus Guyot
19
in a water depth of 1957.6 m. Hole U1372A reached 232.9 meters below sea floor
20
(mbsf) and penetrated 187.3 m of igneous basement below 45.6 m of sediment cover
21
[Koppers et al., 2012a]. The drilled igneous section can be broadly divided into a
22
lower part (104 m thick) of aphyric to variably plagioclase and olivine phyric
23
volcaniclastic breccias succeeded by aphyric to olivine phyric lava flows. Flows in the
24
lower 28.7 m of the upper part of the succession have scoriaceous and oxidized tops,
25
while those above have peperetic tops, implying subaerial to shallow marine
26
conditions.
27
28
1.2 Rigil Guyot (Site U1374):
29
Site U1374 is near the western rift zone on Rigil Guyot in a water depth of 1559 m.
30
Hole U1374A is the deepest hole drilled during IODP Expedition 330, reaching 522
31
mbsf, of which 505.3 m is igneous basement. This consists largely of volcaniclastic
32
breccia, with lava flows increasing in frequency towards the top [Koppers et al.,
33
2012a]. This, together with peperitic tops and bottoms on some of the flow units and
34
sedimentary intervals, suggests that this is an emerging sequence. Aphyric basalt
35
sheets intrude the lower 185.9 m of the hole. Here the phenocryst assemblage is
36
plagioclase-dominated, but olivine becomes more dominant in the upper 256.8 m.
37
Clasts in the lower part of the volcaniclastic breccia have lobate and intricate margins
38
suggesting in-situ cooling, whereas higher up the breccia is more blocky and includes
39
pillow fragments that have broken quenched margins suggesting post-cooling re-
40
deposition [Koppers et al., 2012a].
41
42
1.3 Burton Guyot (Site U1376):
43
Site U1376 is located between two small topographic highs on the eastern side of the
44
summit of Burton Guyot. Drilling reached 182.8 mbsf, penetrating 140.9 m of igneous
45
basement and 41.9 m of sediment cover that included volcanic sand and breccia that
46
provides evidence for late-stage post-erosional volcanism [Koppers et al., 2012a]. The
47
igneous basement consists of a lower 70.2 m of olivine phyric and aphyric
48
volcaniclastic breccia intruded by sheets of aphyric basalt. The upper 17.35 m
49
contains abundant glassy material. The top of the volcaniclastic breccia is cut by an
50
erosion surface, above which is a 33 m thick massive olivine-augite flow. This is
51
succeeded by further breccia and pillow lavas, suggesting that the basement section
52
was emplaced in a submarine environment.
53
54
1.4 Hadar Guyot (Site U1377):
55
Site U1377 is located at the center of the Hadar’s summit plain. Hole U1377B
56
reached 37.0 mbsf, of which 27.9 m was igneous basement. The lower 12.8 m of the
57
igneous sequence consists of what were believed to be a stack of pillow lavas
58
separated by thick well preserved glassy margins [Koppers et al., 2012a]. However,
59
the margins include some curious ‘intrusive’ features where the glass connects to the
60
more massive interior of the unit below. This was interpreted to be where still molten
61
pillow interiors had broken out into space between overlying pillow bodies or where
62
magma had injected into a stack of pillows [Koppers et al., 2012a].
63
2. Analytical methods
64
2.1 Major elements
65
All major element analyses were carried out on fresh glass fragments that were
66
handpicked to avoid any visible alteration. Major element data for each sample are the
67
average of ten spots on a single glass chip, and were measured using a JEOL JXA-
68
8200 Superprobe electron microprobe at the GeoZentrum Nordbayern, Universität
69
Erlangen-Nürnberg. An acceleration voltage of 15 kV, a beam current of 15 nA, and a
70
defocused beam (10 µm) were used. Counting times were 20 s for peaks and 10 s for
71
backgrounds for SiO2, TiO2, Al2O3, FeO, MnO, MgO, CaO, K2O, Na2O, P2O5 and S,
72
while for F and Cl counting times on the peaks and backgrounds were increased to 40
73
and 20 s, respectively. The method used by Brandl et al. [2012] was followed except
74
data were not normalized to glass standards VG-2 and VG-A99. Glass standards VG-
75
2, VG-A99 and VG-568 were analyzed periodically as unknowns to ensure there was
76
no spectrometer drift.
77
78
2.2 Trace elements
79
The trace elements were analyzed in-situ on the same glasses that had previously been
80
analyzed for major elements using an UP193FX New Wave Research Laser operated
81
on an Agilent 7500i ICP-MS at the GeoZentrum Nordbayern, Universität Erlangen-
82
Nürnberg. An argon-helium mixture was used as carrier gas, 25µm spots were
83
analyzed with a 20Hz repetition rate, 0.67 GW/cm2 irradiance and 3.4 J/cm2 fluence.
84
Maximum peak measuring times were 10 msec for 7Li, 29Si, 55Mn; 25 m sec for 45Sc,
85
51
86
140
Ce, 141Pr, 146Nd, 147Sm, 153Eu, 157Gd, 159Tb, 163Dy, 165Ho, 166Er, 169Tm, 172Yb, 175Lu,
87
178
Hf,
88
NIST SRM 612 was used [Pearce et al., 1997]. To ensure accuracy and
V, 53Cr, 59Co, 60Ni, 63Cu, 66Zn, 85Rb, 88Sr, 89Y, 90Zr, 93Nb, 118Sn, 133Cs, 137Ba, 139La,
182
W,
185
Re,
208
Pb,
232
Th,
238
U; and 30 msec for
181
Ta. For external calibration
89
reproducibility NIST SRM 614 and BCR-2G were periodically analyzed (see
90
auxiliary Table S2). Silica contents from the microprobe analyses were used as an
91
internal standard.
92
93
2.3 H2O and CO2 measurements
94
Fifty-two of the 113 samples analyzed for major and trace elements were successfully
95
analyzed for H2O and CO2 using micro-Fourier-transform infrared (FTIR)
96
spectroscopy at the Institute for Research for Earth Evolution (IFREE), Japan Agency
97
for Marine Earth Science and Technology (JAMSTEC). For 34 of these samples,
98
glassy chips were prepared as wafers, analyzed by micro-FTIR spectroscopy, and then
99
the same wafers were mounted and analyzed by EPMA and finally LA-ICP-MS (see
100
Table S1). For the other 18 samples a glassy chip was broken off from the sample
101
prepared for EPMA and LA-ICP-MS and prepared separately for micro-FTIR
102
spectroscopy. In all cases chips were prepared for micro-FTIR spectroscopy by
103
mounting in Orthocryl® and then grinding down one side with silica carbide paper
104
until a large (at least 1 mm across) flat area was exposed. This was then polished
105
using diamond suspension to 1 µm. After polishing, the sample, still within
106
Orthocryl®, was mounted polished side down on a glass slide using Crystal Bond
107
509. The Orthocryl® sample mount was then cut to a thickness of approximately 1
108
mm using a Struers Discoplan-TS saw, before grinding to approximately 150 – 200
109
µm thickness using silica carbide paper, and polishing the second surface, parallel to
110
the first, to 1 µm using diamond suspension. The Crystal Bond 509 and Orthocryl®
111
were then dissolved in acetone, leaving a free-standing wafer approximately 150 –
112
200 µm in thickness with two parallel polished surfaces.
113
114
These wafers were analyzed using a Varian FTS 7000 spectrometer and an attached
115
UMA600 microscope. The spectra used to measure H2O and CO2 were collected
116
conventionally in transmitted light at single spots 20  20 µm square, selected using
117
the microscope, over 512 scans at a resolution of 8 cm-1 using a heated ceramic
118
(globar) infra-red source and a Ge-coated KBr beamsplitter. The wafers were placed
119
on a H2O-free IR-invisible KBr window. Background analyses were taken through the
120
window, before the wafer of glass was positioned in the beam path to measure sample
121
spectra. H2O concentrations were calculated using a modified Beer-Lambert law
122
[Stolper, 1982], with the absorbance at ~3550 cm-1 used for total H2O and that at
123
~1660 cm-1 for molecular H2O (H2Omol) and molar absorptivity coefficients of 63 ± 5
124
and 25 ± 3 l/molcm, respectively [Dixon et al., 1988]. The absorbance was defined as
125
the maximum height of the peaks above a linear baseline. The concentration of OH
126
was obtained by subtracting the concentration of H2Omol from total H2O. The doublet
127
at ~1515 and ~1435 cm-1, indicative of CO2 dissolved in basaltic glass [Fine and
128
Stolper, 1985], was not detected in any of the Louisville glasses. Glass densities, for
129
use in the Beer-Lambert law, were calculated from the oxide compositions of the
130
glasses using the principles, partial molar volumes and volume changes with
131
temperature from Lange and Carmichael [1987] and Lange [1997], including those
132
for H2O from Ochs and Lange [1999]. Thickness for most analyses was obtained from
133
interference fringes on FTIR spectra measured in reflected light on exactly the same
134
spot as those measured in transmitted light following the procedures of Wysoczanski
135
and Tani [2006] and Nichols and Wysoczanski [2007] and using a refractive index of
136
1.546 for basalt [Kumagai and Kaneoka, 2003]. In some cases thicknesses were also
137
measured directly with a digital displacement gauge and there is good agreement
138
between the two methods. Where possible, thicknesses measured using interference
139
fringes were preferred because these were measured in exactly the same spot as the
140
spectra in transmitted light. Uncertainties in H2O species concentrations are estimated
141
to be ±10 % [Wysoczanski and Tani, 2006], which is mostly due to the errors in the
142
thickness and molar absorptivity coefficients. The minimum possible detectable peak
143
on the FTIR spectra has an absorbance of 0.01 (ca. three times background), with the
144
detection limits for each H2O and CO2 species dependent on the thickness at each
145
analyzed spot.
146
147
2.4 Oxygen isotope analyses
148
Oxygen isotope (δ18O) values were analyzed by laser fluorination using a 25 W-
149
Synrad CO2-laser and F2 as reagent at the GeoZentrum Nordbayern, Universität
150
Erlangen-Nürnberg using the method described in detail by Haase et al. [2011] and
151
Genske et al. [2013]. The oxygen isotopes were analyzed on a ThermoFisher Delta
152
Plus mass spectrometer at the GeoZentrum Nordbayern. During each measurement
153
day, four standard samples (UWG-2, NBS-30) were processed and measured together
154
with the samples. The δ18O raw values of a run were corrected by the mean difference
155
of the reference values from the standards (5.8 and 5.1‰, respectively). All oxygen
156
isotope values are given in permil relative to V-SMOW. Reproducibility of the UW
157
GMG 2 garnet standard obtained during the run of this study is 5.84 ± 0.07 ‰ (1sd, n
158
= 29).
159
160
3 Estimating subsidence
161
In order to estimate the subsidence experienced by each sample, the collection depth
162
(depth in drill hole plus water depth), corrected for sediment loading, is subtracted
163
from the paleo-quenching depth derived from the volatile saturation pressures, after
164
accounting for the change in sea level since the formation of the seamount to the
165
present day. For each hole, sediment loading (ds) is calculated using the equation of
166
Crough [1983]:
167
d s = ts *
168
where ts is the thickness of the overlying sediment (m) from Koppers et al., [2012a],
169
m the density of the mantle (3,300 kgm-3), s the density of the sediment (estimated
170
to be 1,900 kgm-3) and w the density of seawater (1,030 kgm-3). This is then
171
subtracted from each sample depth to correct for sediment loading. Sea level at the
172
time the seamounts were erupting is estimated using the curve of Miller et al. [2005],
173
which shows that it has fallen ~44, ~43, ~43 and ~62 m since Canopus (74 Ma), Rigil
174
(69.5 Ma), Burton (66 Ma) and Hadar (50 Ma) formed, respectively.
( r m - rs )
(r m - r w )
[1]
175
176
Additional references:
177
Crough, S. T. (1983), The Correction for Sediment Loading on the Seafloor, J.
178
Geophys. Res., 88(B8), 6449-6454, doi: 10.1029/JB088iB08p06449.
179
Dixon, J. E., E. Stolper, and J. R. Delaney (1988), Infrared spectroscopic
180
measurements of CO2 and H2O in Juan de Fuca Ridge basaltic glasses, Earth Planet.
181
Sci. Lett., 90, 87-104, doi:10.1016/0012-821X(88)90114-8.
182
Fine, G., and E. Stolper (1985), Dissolved carbon dioxide in basaltic glasses:
183
concentration and speciation, Earth Planet. Sci. Lett., 76, 263-278, doi:10.1016/0012-
184
821X(86)90078-6.
185
Genske, F. S., C. Beier, K. M. Haase, S. P. Turner, S. Krumm, and P. A. Brandl
186
(2013), Oxygen isotopes in the Azores islands: crustal assimilation recorded in
187
olivine, Geology, 41, 491-494, doi: 10.1130/G33911.1.
188
Haase, K. M., S. Krumm, M. Regelous, and M. Joachimski (2011), Oxygen isotope
189
evidence for the formation of silicic Kermadec island arc and Havre–Lau backarc
190
magmas by fractional crystallisation, Earth Planet. Sci. Lett., 309, 348-355, doi:
191
10.1016/j.epsl.2011.07.014.
192
Kumagai, H., and I. Kaneoka (2003), Relationship between submarine MORB glass
193
textures and atmospheric component of MORBs, Chem. Geol., 200, 1-24, doi:
194
10.1016/S0009-2541(03)00077-9.
195
Lange, R. A. (1997), A revised model for the density and thermal expansivity of K2O-
196
Na2O-CaO-MgO-Al2O3-SiO2 liquids from 700 to 1900 K: extension to crustal
197
magmatic
198
10.1007/s004100050345.
199
Lange, R. A., and I. S. E. Carmichael (1987), Densities of Na2O-K2O-CaO-MgO-
200
FeO-Fe2O3-Al2O3-TiO2-SiO2 liquids: New measurements and derived partial molar
201
properties,
202
7037(87)90368-1.
203
Miller, K. G., M. A. Kominz, J. V. Browning, J. D. Wright, G. S. Mountain, M. E.
204
Katz, P. J. Sugarman, B. S. Cramer, N. Christie-Black, and S. F. Pekar (2005), The
205
Phanerozoic Record of Global Sea-Level Change, Science, 310, 1293-1298, doi:
206
10.1126/science.1116412.
207
Nichols, A. R. L., and R. J. Wysoczanski (2007), Using micro-FTIR spectroscopy to
208
measure volatile contents in small and unexposed inclusions hosted in olivine crystals,
209
Chem. Geol., 242, 371-384, doi:10.1016/j.chemgeo.2007.04.007.
210
Ochs, F. A., and R. A. Lange (1999), The density of hydrous magmatic liquids,
211
Science, 283, 1314-1317, doi: 10.1126/science.283.5406.1314.
212
Pearce, N. J. G., W. T. Perkins, J. A. Westgate, M. P. Gorton, S. E. Jackson, C. R.
213
Neal, and S. P. Chenery (1997), A compilation of new and published major and trace
214
element data for NIST SRM 610 and NIST SRM 612 glass reference materials,
215
Geostand. Newslett., 21, 115-144, doi: 10.1111/j.1751-908X.1997.tb00538.x.
temperatures,
Geochim.
Contrib.
Cosmchim.
Mineral.
Acta,
51,
Petrol.,
2931-2946,
130,
1-11,
doi:
doi:10.1016/0016-
216
Stolper, E. (1982), Water in silicate glasses: An infrared spectroscopic study, Contrib.
217
Mineral. Petrol., 81, 1-17, doi: 10.1007/BF00371154.
218
Wysoczanski, R. J., and K. Tani (2006), Spectroscopic FTIR imaging of water species
219
in silicic volcanic glasses and melt inclusions: An example from the Izu-Bonin arc, J.
220
Volcanol. Geotherm. Res., 156, 302-314, doi:10.1016/j.jvolgeores.2006.03.024.
Download