G31158-Evans-exstyled

advertisement
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
1
Lizardite versus antigorite serpentinite: Magnetite,
2
hydrogen, and life(?)
3
Bernard W. Evans
4
Department of Earth and Space Sciences, Box 351310, University of Washington, Seattle,
5
Washington 98195-1310, USA
6
ABSTRACT
7
The serpentinization of peridotite operates according to one or other, or a
8
combination, of two end-member mechanisms. In low-temperature environments (50–
9
300 °C), where lizardite is the predominant serpentine mineral, olivine is consumed by
10
reaction with H2O but its composition (Mg#) remains unchanged. Mg-rich lizardite,
11
magnetite and dihydrogen gas (±brucite) are products of the reaction. At higher
12
temperatures (400–600 °C) rates of MgFe diffusion in olivine are orders of magnitude
13
faster, with the result that the growth of Mg-rich antigorite can be accommodated by a
14
compositional adjustment of olivine, eliminating the need to precipitate magnetite and
15
evolve hydrogen. This latter end-member mechanism probably best reflects the situation
16
in the forearc mantle wedge.
17
INTRODUCTION
18
The serpentinization reaction is ordinarily written as conserving oxygen. Because
19
it is accompanied by the oxidation of iron, the mass balance necessarily involves the
20
liberation of dihydrogen gas. The precipitation of magnetite, uptake of ferric iron in
21
lizardite (Whittaker and Wicks, 1970; Seyfried et al., 2007; Evans, 2008), and evolution
22
of hydrogen (Thayer, 1966; Barnes et al., 1972; Neal and Stanger, 1983; Abrajano et al.,
Page 1 of 19
23
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
1990; Sherwood-Lollar et al., 1993) are features widely observed as accompanying the
24
low-temperature serpentinization of peridotite. So for dunite the generalized equation is:
25
MgFe-olivine + H2O = MgFe-serpentine + MgFe-brucite + magnetite +
26
H2(aq) . (1)
27
The inverse correlation between magnetic susceptibility and density that Equation
28
1 implies has been confirmed by measurements (Coleman, 1971; Toft et al., 1990; Oufi et
29
al., 2002). Together with seismic velocities, these properties constitute valuable tools for
30
geophysicists assessing the percentage of serpentinite in subducted and underthrust
31
(wedge) mantle. Equally important, the evolution of hydrogen has been recognized as a
32
source of energy for the potential production of abiotic hydrocarbons via Fischer-Tropsch
33
reactions.
To the author’s knowledge, no serious attempt has been made to explain why the
34
35
low-T serpentinization reaction should be an oxidizing one.
36
Textbook metamorphic reactions involving solid solution minerals (e.g., Spear,
37
1993) are generally written oxygen conserved and unaccompanied by oxidation of iron.
38
For a long time it has been axiomatic in metamorphic petrology that reactions among
39
solid-solution minerals tend to proceed so as to maintain the equivalence of the chemical
40
potentials of components among the phases (in cation exchanges such as FeMg–1, Fe3+Al–
41
1,
42
according to this equilibrium principle. Coexisting solid solution minerals are represented
43
in projection by intersecting T –X or P – X phase loops, and the loci of points on the loops
44
conform to the exchange coefficient KD. Progress of such reactions, whether prograde or
45
retrograde, requires the free exchange of components among the minerals.
KNa–1). PTX-relations, including pseudosections, are routinely calculated and graphed
Page 2 of 19
46
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Ferromagnesian minerals, as in Barrovian metamorphism, are prime examples. Changes
47
in the redox state of iron are not predicted. It seems to have escaped notice that the
48
reaction describing the low-temperature formation of serpentine from peridotite does not
49
follow this classic metamorphic model. Perhaps, quite reasonably, we have simply not
50
had that expectation.
51
At this point it is useful to draw a distinction between lizardite serpentinization,
52
which is generally inferred to occur in the range 50–300 °C, and antigorite
53
serpentinization in the range 400–600 °C. In both cases we can expect the first-formed
54
layer of serpentine to be Mg-rich, a reflection of Fe/Mg partition between olivine (or
55
orthopyroxene) and serpentine (KD ~0.4) and the chemical potentials of Fe and Mg (more
56
specifically the Fe2+Mg-1 exchange potential) imposed on the thin film of fluid at the
57
growth surface (Evans, 2008). Contrary to a seemingly popular view, the Fe-content of
58
the serpentine is not limited by an intrinsic instability of more Fe-rich serpentine, as any
59
global compilation of lizardite analyses can testify (e.g., Evans et al., 2009).
60
THE DIFFUSION PROBLEM
61
Next we should question whether the reactant olivine truly “reacts” in the sense of
62
changing its Fe/Mg ratio in accordance with reaction progress, or whether it merely
63
dissolves at constant composition? Informative in this context are laboratory data on Fe-
64
Mg diffusion in olivine, which is by far the slowest step in a full serpentinization process.
65
At 100 MPa and 200 °C, DFeMg after extrapolation is –32 log units (m2/s) along [001], the
66
fastest direction, and the characteristic diffusion distance (Dt) over 1 my is 104 micron
67
(Dohmen and Chakraborty, 2007). At 500 and 600 °C and the same pressure, DFeMg is
68
10-23.6 and 10-21.9 (m2/s) and the characteristic diffusion distances over 1 my are
Page 3 of 19
69
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
respectively 8 and 60 microns (Fig. 1). Thus, in antigorite serpentinization, especially
70
above 500 °C and over several my, we might expect reaction behavior to follow the
71
textbook metamorphic model, there being no serious limitation on the export of Mg to the
72
growing serpentine and consequent enrichment in iron of the olivine. As a result, there
73
may be little or no growth of magnetite, possibly more brucite, and no yield of hydrogen.
74
In lizardite serpentinization, olivine dissolves but it cannot react by changing its
75
composition. In this case the precipitation of magnetite and evolution of hydrogen are
76
assured. The system as a whole does not attain a condition of minimum Gibbs free energy
77
because the mass-balance contribution of olivine interiors is not included.
78
These rates of Fe-Mg exchange equilibration stand in marked contrast to the rate
79
of serpentinization of olivine as measured hydrothermally in the laboratory (Martin and
80
Fyfe, 1970; Wegner and Ernst, 1983). For example, at 310 °C and 100 MPa, 50 °C below
81
the reaction curve, the hydration reaction is 50% complete after only 20 days. Thus, even
82
at the low temperatures of lizardite serpentinization, the rate of the net-transfer reaction
83
will be limited only by the rate of supply of H2O (MacDonald and Fyfe, 1985). At the
84
higher temperatures of antigorite serpentinization, if hydration is facilitated by a copious
85
supply of H2O, the time for diffusion may be curtailed, and reaction will behave more
86
like lizardite serpentinization.
87
LIZARDITE
88
Consistent with the foregoing, microprobe analyses of low-T serpentinites show
89
that the growth of Mg-rich lizardite from olivine is not accompanied by a corresponding
90
enrichment in its fayalite content. The most frequent composition of lizardite in
91
peridotites of mantle origin has an Mg# of 0.97–0.96 (Evans et al., 2009, Figure 1). In
Page 4 of 19
92
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
many instances though, the Fe content is higher owing to its uptake of ferric iron. These
93
compositions of lizardite reflect an attempt to equilibrate with the Fe2+Mg -1 exchange
94
potentials of the primary olivine and orthopyroxene. Although olivine incrementally
95
dissolves in the process, it is “non-reactive” in the sense of being unable to adjust its
96
composition. For mass-balance another Fe mineral must form, such as magnetite or
97
awaruite Ni3Fe. Although brucite tends to be iron rich, for example with Mg#s of 0.85–
98
0.70, it is not present in sufficient quantities to satisfy the Fe mass-balance. Lizardite
99
serpentinization may be viewed then as a “factory” that adds H2O to a semi-infinite
100
supply of olivine to continuously yield products consisting of Mg-rich lizardite,
101
magnetite, hydrogen, and in some cases brucite and FeNiCo alloys. Thus the
102
mineralogical properties of most serpentinites are the consequence of the enormous
103
differences in effective rates between the hydration reaction and MgFe-diffusion in
104
reactant olivine.
105
ANTIGORITE
106
The growth of antigorite at the expense of olivine will reflect progress of one or
107
more of four possible hydration reactions in the MSH system above ~300 °C: Fo + H2O
108
 Atg + Brc, Fo + Tlc + H2O  Atg, Fo + En + H2O  Atg, and Fo + SiO2(aq) + H2O
109
 Atg. In the CMSH system, antigorite can form along with diopside: Fo + Tr + H2O 
110
Atg + Di, as inferred for the Happo ultramafic complex, Japan, believed to be mantle
111
wedge (Nozaka, 2005). In the FMSH system, as argued above, the higher the temperature
112
the less likely will be an accompanying precipitation of magnetite. These reactions could
113
take place in a subducting slab if they are initiated by the introduction of H2O into
114
anhydrous peridotite, much as anhydrous basalt is converted into a greenschist or
Page 5 of 19
115
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
subgreenschist facies assemblage. A likely alternative in the slab is antigorite growth at
116
the expense of early-formed, oceanic lizardite and chrysotile, forming prograde antigorite
117
serpentinite (O’Hanley, 1996). In this case, the antigorite serpentinite will likely inherit
118
some magnetite. On the other hand, if antigorite grows directly from anhydrous FMSH
119
peridotite at temperatures as high as 500–600 °C, we would predict little or no magnetite
120
nor evolution of hydrogen. These conditions of alteration are to be expected in the mantle
121
wedge above subducted ocean crust that is yielding aqueous fluid because of thermally
122
driven dehydration reactions.
123
There is in fact evidence from examples of antigorite serpentinization, notably
124
from the mantle wedge, that the growth of Mg-rich antigorite from MgFe-olivine results,
125
as anticipated, in Fe-enrichment of the olivine. Amounts of magnetite are small or even
126
zero. The potential for magnetite and serpentine growth is in any case less than in
127
lizardite serpentinization because olivine is stable above ~400 °C. The modal amount of
128
antigorite is constrained by the whole-rock composition: unless silica is introduced, there
129
will be no antigorite in a meta-dunite, and in a harzburgite (e.g., 80% olivine, 20%
130
enstatite) there will be less than 50% antigorite. Estimates of the degree of
131
serpentinization in the mantle wedge based on seismic velocities range widely: 15% to
132
60% (Zhao et al., 2001; Bostock et al., 2002; Courtier et al., 2004).
133
Smith (2010) described an antigorite-rich metaperidotite (sample N15-GN) from
134
the Green Knobs diatreme in the Navajo Volcanic Field that he inferred had formed by
135
hydration of dunite above the Farallon plate. Modally the sample consists of 33% olivine
136
(Mg# 0.884), 63% antigorite (Mg# 0.942), 0.5% diopside, 0.3% magnesite, 0.1%
137
chromite, and 0.1% clinohumite. No magnetite is present. The whole-rock composition
Page 6 of 19
138
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
reconstructed from mineral and modal analyses has Mg# = 0.918, which suggests that the
139
protolith olivine was in the range 0.92- 0.89 that is typical for unaltered Colorado Plateau
140
xenoliths. “The olivine is more Fe-rich than that in any other Green Knobs
141
peridotites…the Fe-rich composition may be attributed to formation of antigorite and
142
partition of Fe into residual olivine” (Smith, 2010).
143
In Mariana forearc serpentinite seamounts, Murata et al. (2009a,b) found that
144
iron-rich olivine (0.86–0.88) formed at the margins of more Mg-rich, primary mantle
145
olivine (0.89–0.93) in close spatial relationship to antigorite formation – with “only small
146
amounts of magnetite”. Iron-enriched olivine also formed preferentially in stripes along
147
(100) olivine “cleavages” (deformation zones). MnO is also enriched in olivine in the
148
stripes, consistent with its partition with respect to antigorite (Trommsdorff and Evans,
149
1974). Formation temperatures were estimated at 450–550 °C. These samples provide
150
good evidence for the influence of MgFe pipe-diffusion along planar defects associated
151
with dislocations in mantle olivine (Ando et al., 2010), perhaps aided by ingress of water
152
(Boudier et al., 2009).
153
Forearc antigorite serpentinites of Neoproterozoic age from the Eastern Desert,
154
Egypt (Khalil and Azer, 2007) contain highly strained olivine (Mg# > 0.89) of mantle
155
derivation, and later unstrained olivine (0.84–0.85). The authors suggest a thermal event
156
(granite intrusion) for the latter; nevertheless, whole-rock data show that the unstrained
157
olivines grew in peridotite with an Mg# of 0.91.
158
In samples of meta-peridotite from serpentinite mélange in Guatemala, antigorite-
159
olivine serpentinite is reported in some cases to be free of magnetite (G. Harlow, personal
160
commun., February, 2010). In two cases, two generations of olivine were found: Mg#s
Page 7 of 19
161
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
0.90 and Mn-enriched 0.84 in one, and 0.91 and Mn-enriched 0.85 in the other. The
162
absence of any indication of pre-existing lizardite, and association with sediment-
163
fertilized HP blocks, supports a supra-subducted plate origin.
164
Antigorite formed from olivine in the supra-subduction zone Trinity peridotite,
165
California, above and close to the Trinity thrust (Peacock, 1987). Although magnetite is
166
present along fractures in the olivine, antigorite growth (as blades penetrating the olivine)
167
was interpreted as largely postdating this magnetite. Other minerals are tremolite,
168
chlorite, talc, carbonate, and ferrit-chromite. The antigorite serpentinization was caused
169
by infiltrating fluids at ~500 °C and ~5 kbar derived from the subducting slab beneath.
170
Field examples of the hydration of olivine directly to antigorite are far fewer than
171
in the case of lizardite. Interpenetrating textures are typical (Wicks and Whittaker, 1977;
172
O’Hanley, 1996). However, it is not clear from their examples how much magnetite was
173
produced and whether or not the olivine underwent an increase in iron. In a good many
174
serpentinites, there is textural evidence for the growth of antigorite at the expense of
175
lizardite and chrysotile. The MSH phase diagram (Evans, 2004) suggests temperatures in
176
excess of 300 °Cfor the growth of antigorite from lizardite.
177
Texturally well-equilibrated metamorphic antigorite-olivine serpentinites
178
ordinarily show the attainment of FeMg-exchange equilibrium between olivine and
179
antigorite (Fig. 2). Excellent examples come from regionally metamorphosed Tethyan
180
ophiolites in the Alps, which reached temperatures on the order of 400–550 °C or more
181
over a metamorphic cycle of perhaps 10 my. Temperatures were evidently sufficiently
182
high for olivine, either by recrystallization or solid-state diffusion, to adjust to the
183
demands of the partition coefficient KD and modal mineralogy. These conditions favor
Page 8 of 19
184
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
the textbook hydration model for serpentinization in which iron is not oxidized. In some
185
cases however, a prograde metamorphic path via lizardite serpentinite is recognizable:
186
olivine Mg#s can be as high as 0.97–0.94, magnetite is abundant, and bodies of rodingite
187
and ophicarbonate occur. In other samples (Trommsdorff and Evans, 1972; Worden et al.,
188
1991; Scambelluri et al., 1991), olivine falls in the range 0.89–0.84, while whole-rock
189
Mg#s are around 0.91–0.90. This suggests recrystallized olivine in FeMg-exchange
190
equilibrium with antigorite, and minimal prior oxidation. These samples do contain ~1%
191
spinel but it is generally in the ferrit-chromite range (Evans and Frost, 1975), sometimes
192
in aggregates ~0.1 mm in diameter suggesting derivation by alteration of primary
193
chromite.
194
EXPERIMENTS AND MODELING
195
Hydrothermal laboratory experiments on MgFe-olivine at low temperature have
196
produced magnetite along with lizardite, and presumably some hydrogen (Moody, 1976;
197
Janecky and Seyfried, 1986; Normand et al., 2002; Allen and Seyfried, 2003). Even at
198
500 or 600 °C the characteristic MgFe diffusion distance in olivine over one year is so
199
small that laboratory experiments on antigorite growth from MgFe-olivine are equally
200
unlikely to follow the textbook model of reaction progress, and so we should expect them
201
to produce magnetite. Conversely, at very low temperatures, e.g., 100–200 °C, it seems
202
that magnetite formation is somewhat diminished because of the take-up of ferric iron in
203
lizardite (Seyfried et al., 2007; Evans, 2008; Klein et al., 2009).
204
Reaction-path modeling that assumes a constant composition of olivine
205
(McCollom and Bach, 2009; Klein et al., 2009) is entirely appropriate for low-
206
temperature serpentinization in nature; the “full equilibrium” model would allow reaction
Page 9 of 19
207
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
progress to enrich Ol, Srp and Brc in iron. However, model W/R ratios as high as 1:1
208
probably overestimate the amounts of magnetite, antigorite, and H2 that result
209
(irrespective of MgFe-diffusion in olivine) from internal oxygen and silica buffering
210
down-T along the isobaric MFSH quasi-univariant curve Ol-Atg-Mag. If, as is likely in
211
nature, fluid-absent conditions prevail during the decompression of mantle peridotite, the
212
oxygen and silica buffer capacities are likely to endure during cooling until olivine finally
213
succumbs to a major influx of fluid and low-T serpentinization ensues. Therefore, we can
214
largely rule out much magnetite formation and hydrogen production during cooling prior
215
to the major serpentinizing event, which ordinarily is related to faulting of some kind.
216
At the isobaric MFSH quasi-invariant point Ol + Liz + Brc + Mag in dunite
217
(reaction 1), the loss or armoring of olivine caused by hydration releases a buffer
218
constraint that in turn allows ferroan brucite to accept silica from any source and be
219
converted to serpentine, magnetite, H2 and H2O (e.g., Bach et al., 2006; Frost and Beard,
220
2007; Beard et al., 2009). Large inputs of ocean or ground water will eventually
221
overwhelm the oxygen-buffer capacity of the rock and yield more magnetite (or hematite)
222
and hydrogen by oxidation of olivine, serpentine and brucite.
223
ABIOGENIC METHANE
224
The hydrogen-rich fluids found issuing from serpentinite on land and on the ocean
225
floor are often also enriched in methane. As many have pointed out, when hydrogen can
226
be combined with CO2, potential abiotic hydrocarbons are possible via a natural
227
equivalent of Fischer-Tropsch synthesis. Since the discovery of the serpentinite-hosted
228
Lost City hydrothermal field, 15 km W of the Mid-Atlantic Ridge (Kelley et al., 2001;
229
2005; Früh-Green et al., 2003), many workers, for example, Sleep et al. (2004), Schulte
Page 10 of 19
230
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
et al. (2006), Martin et al. (2008), Proskurowski et al. (2008), have seriously advanced
231
the idea of a serpentinite origin for life on the planets. The necessary ingredients
232
(peridotite or komatiite and water) were available together on the surfaces of the early
233
planets.
234
CONCLUDING REMARKS
235
The contrasting mechanisms of lizardite and antigorite serpentinization suggest
236
that the production of H2, and consequently potential abiogenic CH4, may not be an
237
accompaniment of mantle wedge serpentinization. Similarly, the amount of magnetite
238
generated in mantle wedge serpentinite may be much less than found in ocean floor
239
serpentinites. Therefore the magnetic anomalies these rocks induce (e.g., Hyndman and
240
Peacock, 2003; Blakely et al., 2005) may be quite different. There are certainly risks in
241
taking serpentinized abyssal peridotites as models for the kind of serpentinite that occurs
242
in the mantle wedge. On the other hand, the model for abiogenic hydrocarbons and
243
potential life on planetary surfaces is secure because of the low temperatures involved. It
244
is a sobering thought that we might owe our existence on earth to the sluggish low-
245
temperature FeMg lattice diffusion in olivine in peridotite.
246
ACKNOWLEDGMENTS
247
I thank N.I. Christensen, R. Dohmen, B.R. Frost, G.E. Harlow, P.B. Kelemen,
248
H. Maekawa, O. Müntener, D. Smith, and F.J. Wicks for helpful comments on the
249
manuscript.
250
REFERENCES CITED
251
Abrajano, T.A., Sturchio, N.C., Kennedy, B.M., Lyon, G.L., Muehlenbachs, K., and
252
Bohlke, J.K., 1990, Geochemistry of reduced gas related to serpentinization of the
Page 11 of 19
253
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Zambales ophiolites, Philippines: Applied Geochemistry, v. 5, p. 625–630, doi:
254
10.1016/0883-2927(90)90060-I.
255
Allen, D.E., and Seyfried, W.E., 2003, Compositional controls on vent fluids from
256
ultramafic-hosted hydrothermal systems at mid-ocean ridges: An experimental study
257
at 400°C, 500 bars: Geochimica et Cosmochimica Acta, v. 67, p. 1531–1542, doi:
258
10.1016/S0016-7037(02)01173-0.
259
Ando, J., Ohfuji, H., Urata, Y., Murata, K., and Maekawa, H., 2010, Heterogeneous
260
distribution of Fe in olivine grains induced by serpentinization: (unpublished
261
manuscript).
262
Bach, W., Paulick, H., Garrido, C.J., Ildefonse, B., Meurer, W.P., and Humphris, S.E.,
263
2006, Unraveling the sequence of serpentinization reactions: petrography, mineral
264
chemistry, and petrophysics of serpentinites from MAR 15°Nn(ODP Leg 209, Site
265
1274): Geophysical Research Letters, v. 33, p. L13306, doi:
266
10.1029/2006GL025681.
267
Barnes, I., Rapp, J.B., O’Neil, J.R., Sheppard, R.A., and Gude, A.J., 1972, Metamorphic
268
assemblages and the direction of fluid flow of metamorphic fluids in four instances
269
of serpentinization: Contributions to Mineralogy and Petrology, v. 35, p. 263–276,
270
doi: 10.1007/BF00371220.
271
Beard, J.S., Frost, B.R., Fryer, P., McCaig, A., Searle, R., Ildefonse, B., Zinin, P., and
272
Sharma, S.K., 2009, Onset and progress of serpentinization and magnetite formation
273
in olivine-rich troctolites from IODP Hole U1309D: Journal of Petrology, v. 50,
274
p. 387–403, doi: 10.1093/petrology/egp004.
Page 12 of 19
275
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Blakely, R.J., Brocher, T.M., and Wells, R.E., 2005, Subduction-zone magnetic
276
anomalies and implications for hydrated forearc mantle: Geology, v. 33, p. 445–448,
277
doi: 10.1130/G21447.1.
278
Bostock, M.G., Hyndman, R.D., Rondenay, S., and Peacock, S.M., 2002, An inverted
279
continental Moho and serpentinization of the forearc mantle: Nature, v. 417, p. 536–
280
538, doi: 10.1038/417536a.
281
Boudier, F., Baronnet, A., and Mainprice, D., 2009, Serpentine mineral replacements of
282
natural olivine and their seismic implications: Oceanic lizardite versus subduction-
283
related antigorite: Journal of Petrology, v. 51, p. 495–512, doi:
284
10.1093/petrology/egp049.
285
Coleman, R.G., 1971, Petrologic and geophysical nature of serpentinites: Geological
286
Society of America Bulletin, v. 82, p. 897–918, doi: 10.1130/0016-
287
7606(1971)82[897:PAGNOS]2.0.CO;2.
288
Courtier, A.M., Hart, D.J., and Christensen, N.I., 2004, Seismic properties of Leg 195
289
serpentinites and their geophysical implications: Proceedings of the Ocean Drilling
290
Program. Scientific Results, v. 195, p. 1–12.
291
Dohmen, R., and Chakraborty, S., 2007, Fe-Mg diffusion in olivine II: point defect
292
chemistry, change of diffusion mechanisms and a model for calculation of diffusion
293
coefficients in natural olivine: Physics and Chemistry of Minerals, v. 34, p. 597–598,
294
doi: 10.1007/s00269-007-0185-3.
295
296
Evans, B.W., 2004, The serpentinite multisystem revisited: Chrysotile is metastable:
International Geology Review, v. 46, p. 479–506, doi: 10.2747/0020-6814.46.6.479.
Page 13 of 19
297
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Evans, B.W., 2008, Control of the products of serpentinization by the Fe2+Mg–1 exchange
298
potential of olivine and orthopyroxene: Journal of Petrology, v. 49, p. 1873–1887,
299
doi: 10.1093/petrology/egn050.
300
Evans, B.W., and Frost, B.R., 1975, Chrome –spinel in progressive metamorphism – a
301
preliminary analysis: Geochimica et Cosmochimica Acta, v. 39, p. 959–972, doi:
302
10.1016/0016-7037(75)90041-1.
303
Evans, B.W., Kuehner, S.M., and Chopelas, A., 2009, Magnetite-free yellow lizardite
304
serpentinization of olivine websterite, Canyon Mountain complex, Oregon: The
305
American Mineralogist, v. 94, p. 1731–1734, doi: 10.2138/am.2009.3301.
306
Frost, B.R., and Beard, J., 2007, On silica activity and serpentinization: Journal of
307
Petrology, v. 48, p. 1351–1368, doi: 10.1093/petrology/egm021.
308
Früh-Green, G.L., Kelley, D.S., Bernasconi, S.M., Karson, J.A., Ludwig, K.A.,
309
Butterfield, D.A., Boschi, C., and Proskurowski, G., 2003, 30,000 years of
310
hydrothermal activity at the lost City Vent Field: Science, v. 301, p. 495–498, doi:
311
10.1126/science.1085582.
312
Hyndman, R.D., and Peacock, S.M., 2003, Serpentinization of the forearc mantle: Earth
313
and Planetary Science Letters, v. 212, p. 417–432, doi: 10.1016/S0012-
314
821X(03)00263-2.
315
Janecky, D.R., and Seyfried, W.E., 1986, Hydrothermal serpentinization of peridotite
316
within the oceanic crust: Experimental investigations of mineralogy and major
317
element chemistry: Geochimica et Cosmochimica Acta, v. 50, p. 1357–1378, doi:
318
10.1016/0016-7037(86)90311-X.
Page 14 of 19
319
320
321
322
323
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Kelley, D.S., et al., 2001, An off-axis hydrothermal vent field near the Mid-Atlantic
Ridge at 30°N: Nature, v. 412, p. 145–149, doi: 10.1038/35084000.
Kelley, D.S., et al., 2005, A serpentinite-hosted ecosystem: the Lost City hydrothermal
field: Science, v. 307, p. 1428–1434, doi: 10.1126/science.1102556.
Khalil, E.E.S., and Azer, M.K., 2007, Supra-subduction affinity in the Neoproterozoic
324
serpentinites in the Eastern Desert, Egypt: Evidence from mineral composition:
325
Journal of African Earth Sciences, v. 49, p. 136–152, doi:
326
10.1016/j.jafrearsci.2007.08.002.
327
Klein, F., Bach, W., Jöns, N., McCollom, T., Moskowitz, B., and Berquo, T., 2009, Iron
328
partitioning and hydrogen generation during serpentinization of abyssal peridotites
329
from 15°N on the Mid-Atlantic Ridge: Geochimica et Cosmochimica Acta, v. 73,
330
p. 6868–6893, doi: 10.1016/j.gca.2009.08.021.
331
MacDonald, A.H., and Fyfe, W.S., 1985, Rates of serpentinization in seafloor
332
environments: Tectonophysics, v. 116, p. 123–135, doi: 10.1016/0040-
333
1951(85)90225-2.
334
Martin, B., and Fyfe, W.S., 1970, Some experimental and theoretical observations on the
335
kinetics of hydration reactions with particular reference to serpentinization:
336
Chemical Geology, v. 6, p. 185–195, doi: 10.1016/0009-2541(70)90018-5.
337
338
339
Martin, W., Baross, J., Kelley, D., and Russell, M.J., 2008, Hydrothermal vents and the
origin of life: Nature Reviews Microbiology, v. 6, p. 805–814.
McCollom, T.M., and Bach, W., 2009, Thermodynamic constraints on hydrogen
340
generation during serpentinization: Geochimica et Cosmochimica Acta, v. 73,
341
p. 856–875, doi: 10.1016/j.gca.2008.10.032.
Page 15 of 19
342
343
344
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Moody, J.B., 1976, An experimental study of the serpentinization of iron-bearing
olivines: Canadian Mineralogist, v. 14, p. 462–478.
Murata, K., Maekawa, H., Yokose, H., Fujioa, K., Ishii, T., Chiba, H., and Wada, Y.,
345
2009a, Significance of serpentinization of wedge mantle peridotites beneath Mariana
346
forearc, western Pacific: Geosphere, v. 5, p. 90–104, doi: 10.1130/GES00213.1.
347
Murata, K., Maekawa, H., Ishii, K., Mohammad, Y.O., and Yokose, H., 2009b, Iron-rich
348
stripe patterns in olivines of serpentinized peridotite from Mariana forearc
349
seamounts, western Pacific: Journal of Mineralogical and Petrological Sciences,
350
v. 104, p. 199–203, doi: 10.2465/jmps.081022h.
351
Neal, C., and Stanger, G., 1983, Hydrogen generation from mantle source rocks in Oman:
352
Earth and Planetary Science Letters, v. 66, p. 315–320, doi: 10.1016/0012-
353
821X(83)90144-9.
354
Normand, C., Williams-Jones, A.E., Martin, R.F., and Vali, H., 2002, Hydrothermal
355
alteration of olivine in a flow through autoclave: Nucleation and growth of
356
serpentine phases: The American Mineralogist, v. 87, p. 1699–1709.
357
Nozaka, T., 2005, Metamorphic history of serpentinite mylonites from the Happo
358
ultramafic complex, central Japan: Journal of Metamorphic Petrology, v. 23, p. 711–
359
723, doi: 10.1111/j.1525-1314.2005.00605.x.
360
361
O’Hanley, D.S., 1996, Serpentinites: Records of tectonic and petrologic history: Oxford
UK, Oxford University Press, 277p.
362
Oufi, O., Cannat, M., and Horen, H., 2002, Magnetic properties of variably serpentinized
363
abyssal peridotites: Journal of Geophysical Research, v. 107, B5, p. 2095–2113, doi:
364
10.1029/2001JB000549.
Page 16 of 19
365
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Peacock, S.M., 1987, Serpentinization and infiltration metasomatism in the Trinity
366
peridotite, Klamath province, northern California: implications for subduction zones:
367
Contributions to Mineralogy and Petrology, v. 95, p. 55–70, doi:
368
10.1007/BF00518030.
369
Proskurowski, G., Lilley, M.D., Seewald, J.S., Frűh-Green, G.L., Olson, E.J., Lupton,
370
J.E., Sylva, S.P., and Kelley, D.S., 2008, Abiogenic hydrocarbon production at Lost
371
City Hydrothermal Field: Science, v. 319, p. 604–607, doi:
372
10.1126/science.1151194.
373
Scambelluri, M., Hoogerduijn Strating, E.H., Piccardo, G.B., Vissers, R.L.M., and
374
Rampone, E., 1991, Alpine olivine- and titanian clinohumite-bearing assemblages in
375
the Erro-Tobbio peridotite (Voltri Massif, NW Italy): Journal of Metamorphic
376
Geology, v. 9, p. 79–92, doi: 10.1111/j.1525-1314.1991.tb00505.x.
377
Schulte, M., Blake, D., Hoehler, T., and McCollom, T.M., 2006, Serpentinization and its
378
implication for life on the early Earth and Mars: Astrobiology, v. 6, p. 364–376, doi:
379
10.1089/ast.2006.6.364.
380
Seyfried, W.E., Jr., Foustoukos, D.I., and Fu, Q., 2007, Redox evolution and mass
381
transfer during serpentinization: An experimental and theoretical study at 200°C,
382
4500 bar with implications for ultramafic-hosted hydrothermal systems: Geochimica
383
et Cosmochimica Acta, v. 71, p. 3872–3880, doi: 10.1016/j.gca.2007.05.015.
384
Sherwood-Lollar, B., Frape, S.K., Weise, S.M., Fritz, P., Maacko, S.A., and Welham,
385
J.A., 1993, Abiogenetic methanogenesis in crystalline rocks: Geochimica et
386
Cosmochimica Acta, v. 57, p. 5087–5097, doi: 10.1016/0016-7037(93)90610-9.
Page 17 of 19
387
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Sleep, N.H., Meibom, A., Fridriksson, T., Coleman, R.G., and Bird, D.K., 2004, H2-
388
fluids from serpentinization: geochemical and biotic implications: Proceedings of the
389
National Academy of Sciences of the United States of America, v. 101, p. 12818–
390
12823, doi: 10.1073/pnas.0405289101.
391
Smith, D., 2010, Antigorite peridotite, metaserpentinite, and other inclusions from
392
diatremes on the Colorado Plateau, southwestern USA: Implications for the mantle
393
wedge during low-angle subduction: Journal of Petrology, In press.
394
395
396
397
398
Spear, F.S., 1993, Metamorphic Phase Equilibria and Pressure-Temperature-Time Paths:
Mineralogical Society of America Monograph, 793 p.
Thayer, T.P., 1966, Serpentinization considered as a constant-volume process: The
American Mineralogist, v. 51, p. 685–710.
Toft, P.B., Arkani-Hamed, J., and Haggerty, S.E., 1990, Petrophysics of retrograde
399
serpentinization: Physics of the Earth and Planetary Interiors, v. 65, p. 137–157, doi:
400
10.1016/0031-9201(90)90082-9.
401
Trommsdorff, V., and Evans, B.W., 1972, Progressive metamorphism of antigorite schist
402
in the Bergell tonalite aureole (Italy): American Journal of Science, v. 272, p. 423–
403
437.
404
Trommsdorff, V., and Evans, B.W., 1974, Alpine metamorphism of peridotitic rocks:
405
Schweizerische Mineralogische und Petrographische Mitteilungen, v. 54, p. 332–
406
352.
407
Wegner, W.W., and Ernst, W.G., 1983, Experimentally determined hydration and
408
dehydration reactions in the system MgO-SiO2-H2O: American Journal of Science,
409
v. 283A, p. 151–180.
Page 18 of 19
410
411
412
413
414
Publisher: GSA
Journal: GEOL: Geology
Article ID: G31158
Whittaker, E.J.W., and Wicks, F.J., 1970, Chemical differences among the serpentine
“polymorphs”: A discussion: The American Mineralogist, v. 55, p. 1025–1047.
Wicks, F.J., and Whittaker, E.J.W., 1977, Serpentine textures and serpentinization:
Canadian Mineralogist, v. 15, p. 446–488.
Worden, R.H., Droop, G.T.R., and Champness, P.E., 1991, The reaction antigorite 
415
olivine + talc + H2O in the Bergell aureole, N. Italy: Mineralogical Magazine, v. 55,
416
p. 367–378, doi: 10.1180/minmag.1991.055.380.07.
417
Zhao, D., Wang, K., Rogers, G.C., and Peacock, S.M., 2001, Tomographic image of low
418
P velocity anomalies above slab in northern Cascadia subduction zone: Earth,
419
Planets, and Space, v. 53, p. 285–293.
420
FIGURE CAPTIONS
421
Figure 1. Characteristic diffusion distance in olivine (Mg# 0.9) at 100 MPa. Extrapolated
422
from Dohmen and Chakraborty (2007).
423
424
Figure 2. Roozeboom plot of iron-magnesium partitioning between antigorite and olivine
425
in texturally equilibrated serpentinites. Data from the literature and unpublished work of
426
the author. Note that roughly 10%–30% of the iron in antigorite is ferric iron. For this
427
reason, the trend of data-points does not pass through the origin.
Page 19 of 19
Download