• Aucun résultat trouvé

Measurement of Volume Change and Mass Transfer During Serpentinization: Insights From the Oman Drilling Project

N/A
N/A
Protected

Academic year: 2021

Partager "Measurement of Volume Change and Mass Transfer During Serpentinization: Insights From the Oman Drilling Project"

Copied!
30
0
0

Texte intégral

(1)

HAL Id: hal-02938748

https://hal.archives-ouvertes.fr/hal-02938748

Submitted on 27 Nov 2020

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

During Serpentinization: Insights From the Oman Drilling Project

Benjamin Malvoisin, Chang Zhang, Othmar Müntener, Lukas Baumgartner, Peter Kelemen

To cite this version:

Benjamin Malvoisin, Chang Zhang, Othmar Müntener, Lukas Baumgartner, Peter Kelemen. Mea-

surement of Volume Change and Mass Transfer During Serpentinization: Insights From the Oman

Drilling Project. Journal of Geophysical Research : Solid Earth, American Geophysical Union, 2020,

125 (5), �10.1029/2019JB018877�. �hal-02938748�

(2)

Measurement of volume change and mass transfer

1

during serpentinisation: insights from the Oman

2

Drilling Project

3

Benjamin Malvoisin 1,2, Chang Zhang 1,3, Othmar M¨untener1, Lukas P.

4

Baumgartner1, Peter B. Kelemen 4,5, Oman Drilling Project Science Party

5

1Institut des Sciences de la Terre, Universit´e de Lausanne, Lausanne, Switzerland.

6 2Universit´e Grenoble Alpes, CNRS, ISTerre, 38000 Grenoble, France.

7 3State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy

8

of Sciences, Beijing 100029, China

9 4Lamont-Doherty Earth Observatory, Columbia University, 61 Route 9W, Palisades, NY 10964, United

10

States of America

11 5Department of Earth and Environmental Sciences, Columbia University, New York, NY 10027, United

12

States of America

13

Key Points:

14

A dunite collected during the Oman Drilling Project records serpentinisation at

15

low temperature.

16

The solid volume increases by more than 50 % during reaction as predicted for

17

a closed system.

18

The primary olivine trace element zoning is preserved during serpentinisation.

19

Corresponding author: Benjamin Malvoisin,benjamin.malvoisin@unil.ch

(3)

Abstract

20

Serpentinisation plays a key role on the evolution of the physico-chemical properties

21

of the mantle lithosphere. The rate of serpentinisation reactions is controlled by the

22

transport of fluid which itself depends on volume change during reaction. Element

23

transfer can strongly modify the magnitude and sign of volume change. Here, we mea-

24

sure solid volume change and element transport perpendicular to a serpentine vein

25

in a serpentinised dunite collected at depth during the Oman Drilling Project. The

26

sample is extensively replaced (extent of reaction > 80 %) by a serpentine/brucite

27

mixture parallel to a main serpentine vein network. The Mg content of serpentine

28

and brucite indicates reaction with a small amount of fluid at temperatures below

29

100 C. Concentrations of fluid-mobile trace elements (Na, Ca, Sr, Rb and Ba) de-

30

crease perpendicular to the main vein. Primary olivine contains parallel platelets of a

31

clinopyroxene/magnetite symplectite. Tomography at the nanoscale reveals that these

32

inclusions do not react during serpentinisation but are cracked and displaced. We use

33

these inert markers to measure a 59 to 74 % positive volume change, that is close to

34

the 52 % expected for reaction in a closed system. Chemical data indicate no change

35

in major element composition during reaction except for the addition of water. The

36

initial olivine zoning in Al, Ti, V, Sc and Cr is still preserved in serpentine and brucite.

37

Serpentinisation can thus be a local replacement process during which the solid volume

38

homogeneously increases at the micrometer scale and the transport of aqueous species

39

is limited.

40

1 Introduction

41

Ultramafic rocks react with water at slow spreading ridges, at ocean-continent

42

transitions, in subduction zones and in ophiolites. This process of ”serpentinisation”

43

raises the interest of the scientific community not only because it is ubiquitous but

44

also because it has a first-order impact in geodynamics and in biology (Kelley et

45

al., 2001; Sleep et al., 2004; Delescluse & Chamot-Rooke, 2008; Wada et al., 2008;

46

M¨untener, 2010; Russell et al., 2010; Reynard, 2013) since it produces serpentine

47

((M g, F e)3Si2O5(OH)4), brucite ((M g, F e)OH2), magnetite (F e3O4) and hydrogen

48

(H2) with physico-chemical properties (i.e. composition, density, magnetism, rheol-

49

ogy and energetic potential) strongly different from the primary olivine and pyroxene

50

(Moody, 1976; Escart´ın et al., 2001; Oufi, 2002; McCollom & Bach, 2009; Malvoisin,

51

Carlut, & Brunet, 2012).

52

Klein et al. (2009, 2013) developed thermodynamic models predicting the abun-

53

dance and composition of the products of serpentinisation as a function of tempera-

54

ture and reactant compositions. They predict that magnetite formation mainly occurs

55

above 200C whereas Fe-brucite and Fe-serpentine are the dominant iron-bearing phase

56

below this temperature (Klein et al., 2014). Experiments on powders provide serpen-

57

tinisation rates that are dependent on temperature, pressure, grain size and protolith

58

composition (Martin & Fyfe, 1970; Wegner & Ernst, 1983; Malvoisin, Brunet, et al.,

59

2012; Andreani et al., 2013). However, these constraints are not sufficient to quanti-

60

tatively model serpentinisation in natural systems. Experiments on rock aggregates

61

indeed revealed orders of magnitude slower kinetics compared to rates measured on

62

powders (Malvoisin & Brunet, 2014; Klein et al., 2015). Water transport thus plays

63

a key role on reaction rate and local thermodynamic equilibrium. Some peridotites

64

interacting with near-surface fluids have undergone rapid cooling. This generated fluid

65

pathways through thermal cracking (Demartin et al., 2004; Boudier et al., 2005) and

66

tectonic deformation (Roum´ejon & Cannat, 2014). Understanding the evolution of

67

these pathways and the generation of new pathways during serpentinisation requires

68

understanding of the couplings between reaction and deformation. Serpentine and

69

brucite have a lower density than olivine and pyroxene. Serpentinisation in a closed

70

system is thus expected to induce an increase in solid volume (ΔVV) of approximately

71

(4)

50 % (Macdonald & Fyfe, 1985; O’Hanley, 1992). This change in volume has two

72

opposite consequences for fluid pathway evolution during serpentinisation. It leads to

73

both decrease and increase in permeability through porosity filling by reaction prod-

74

ucts (Farough et al., 2015; Godard et al., 2013) and stress build-up leading to cracking

75

(Macdonald & Fyfe, 1985; O’Hanley, 1992; B. W. Evans, 2004; Jamtveit et al., 2009),

76

respectively. This latter process of reaction-induced cracking is constrained with ob-

77

servations in natural systems (Coleman & Keith, 1971; Loney et al., 1971; Pl¨umper et

78

al., 2012; Roum´ejon & Cannat, 2014; Malvoisin et al., 2017), thermodynamic theory

79

(Kelemen & Hirth, 2012), thermodynamic models of water rock reaction in natural

80

and simplified systems (Klein et al., 2009; Malvoisin, 2015; de Obeso et al., 2017),

81

physical and numerical models (Rudge et al., 2010; Ulven et al., 2014; Shimizu &

82

Okamoto, 2016; Malvoisin et al., 2017; O. Evans et al., 2018, n.d.) and experiments

83

(Zhu et al., 2016; Zheng et al., 2018; Lafay et al., 2018). Despite the key role of solid

84

volume change during serpentinisation, this parameter is not well constrained since it

85

has not been measured in natural samples yet and it is closely related to mass transfer

86

during reaction.

87

The transport of aqueous species during reaction can modify solid volume change

88

(Thayer, 1966; Gresens, 1967; Carmichael, 1987; Fletcher & Merino, 2001; A. Putnis &

89

Austrheim, 2010). For example, pseudomorphic replacement and microtexture preser-

90

vation are interpreted as evidence for isochoric replacement thanks to mass transfer

91

(Merino & Dewers, 1998; Centrella et al., 2015). Some serpentinised peridotites have

92

a different composition than mantle rocks. In addition to water incorporation, the

93

main change in major elements measured in abyssal serpentinised peridotites consists

94

in a decrease of theM gO/SiO2 ratio by 5 % on average (Snow & Dick, 1995; Niu,

95

2004; Monnier et al., 2006; Malvoisin, 2015; de Obeso & Kelemen, 2018). This de-

96

crease can be explained by Si gain or Mg loss during reaction (Malvoisin, 2015; J¨ons

97

et al., 2017; de Obeso & Kelemen, 2018). Enrichment in fluorine, boron and chlorine

98

are also measured in abyssal peridotites (Thompson & Melson, 1970; Bonatti et al.,

99

1984; Sharp & Barnes, 2004; Vils et al., 2008; Bonifacie et al., 2008; Debret et al.,

100

2014; Anselmi et al., 2014). Previous studies on trace elements in abyssal serpentinised

101

peridotites are limited to the hand-specimen scale without in-situ investigations. The

102

bulk geochemistry shows patterns in rare earth elements (REE) highly related to those

103

of their protoliths. Al, V, Sc and Cr are considered as ’immobile’ during hydrother-

104

mal alteration (Paulick et al., 2006). The high field strength elements (HFSEs) show

105

low solubility and mobility in aqueous fluids during hydrothermal alteration relative

106

to REEs. It was thus proposed to use REE and compatible elements to identify the

107

protoliths of serpentinites and to recover the processes experienced by these protoliths

108

(Deschamps et al., 2013). In contrast, U, Pb and Sr contents are characterized by

109

strong enrichments associated with alteration on the seafloor, and in particular with

110

the precipitation of carbonates (Seifert & Brunotte, 1996; Olivier & Boyet, 2006).

111

These changes in composition can influence the geochemical budgets at large scale

112

(Paulick et al., 2006). However, it is difficult to determine the respective contribu-

113

tion of serpentinisation and alteration on the seafloor to mass transfer and porosity

114

evolution in these samples.

115

A way to focus on the mineralogical processes associated with serpentinisation is

116

to work on samples collected by drilling. Tutolo et al. (2016) measured several percent

117

of porosity in serpentinised peridotites drilled in the Atlantis massif. The observed pore

118

diameters are orders of magnitude smaller than the pores formed during alteration on

119

the seafloor (10 μm (J¨ons et al., 2017)). Nanoscale porosity probably plays a key

120

role for fluid transport during replacement reactions but its mechanism of formation

121

remains poorly understood due to a lack of constraints on volume change and mass

122

transfer during reaction.

123

(5)

We measure here solid volume change associated with serpentinisation in a sam-

124

ple drilled during the International Continental Scientific Drilling Program (ICDP)

125

supported Oman Drilling Project. We couple these results with measurements of ma-

126

jor and trace element composition, thermodynamic modelling, and nanotomography.

127

The combination of these tools provides new constraints for determining the miner-

128

alogical processes at play during serpentinisation.

129

2 Sampling and methods

130

We study a core sample from Oman Drilling Project Hole BA4A (sample BA4A-

131

81-1-1-17), drilled into serpentinised peridotites of the Wadi Tayi massif of the Samail

132

Ophiolite, the largest sub-aerial exposure of oceanic lithosphere on Earth. The Samail

133

Ophiolite was formed by igneous accretion at a fast-spreading ridge and then obducted

134

onto the Arabian Continental margin (Boudier et al., 1988). Serpentinisation may have

135

occurred throughout the history of the ophiolite and is still ongoing (Neal & Stanger,

136

1983; Kelemen & Matter, 2008). The extent of hydration (serpentinisation) of core

137

from Hole BA4A well is generally more than 95 %, with only minor relicts of primary

138

mantle minerals. The sample we study comes from one of the freshest portions of the

139

core at a depth of 215 m. It is a dunite, principally composed of serpentine, brucite

140

and olivine in a mesh texture, which is cross-cut by a serpentine vein surrounded by

141

a reaction front. It thus seems ideal for quantifying mass transfer and volume change

142

during reaction, since it allows for comparison between fully and partly reacted rock

143

domains.

144

The variations of the measured values are given with errors corresponding to

145

their standard deviation. The composition and the microtexture of the samples were

146

first characterized at the University of Lausanne with a field-emission scanning elec-

147

tron microscope (Tescan Mira II LMU) operating at 10 kV and 23 mm working dis-

148

tance and equipped with an Energy Dispersive Spectrometer from Oxford Instruments.

149

Serpentine and brucite minerals were identified by Raman microspectrometry at the

150

University of Lausanne with a Horiba LabRAM HR800 equipped with a 532.1 nm

151

laser.

152

2.1 X-ray microtomography

153

A 8 mm and 23 mm high core is scanned with a Skyscan-1173 at the University

154

of Lausanne. 2200 X-ray images are acquired with a 360 rotation of the samples at a

155

step size of 0.225 with a time of acquisition of 800 ms per frame. The acquisition is

156

performed at 70 kV and 114μA by accumulating 33 frames with a voxel size of 11.07

157

μm. The 3D images reconstructed from the transmission images are segmented and

158

analysed with Blob3D (Ketcham, 2005). We develop a Matlab code to display the

159

result of the image analysis and to extract the extent of reaction from the segmented

160

3D image. The extent of reaction is calculated based on the measured volumes of

161

olivine, spinel and brucite/serpentine in a 350μmwide sliding window parallel to the

162

orientation of the vein cross-cutting the sample. The similar attenuation of olivine

163

and Fe-brucite (identified with Scanning Electron Microscopy) makes segmentation

164

difficult. These phases can be distinguished based on their aspect ratio (e) calculated

165

as the length ratio of the minimum and maximum axes of an ellipsoid fitted to each

166

segmented grain (Figure S1 in Supplementary materials). Fe-brucite has a lowerethan

167

olivine due to its platy crystal form. Therefore, the olivine volume used to calculate

168

the extent of reaction displayed in Figure 1 is determined by only considering grains

169

with e > 0.35. We also exclude from olivine grain segmentation groups of voxels

170

with a volume smaller than 1.5.105 μm3. The radius of a sphere of this latter volume

171

corresponds to three times the voxel size. The extent of reaction estimated with X-

172

(6)

-10 -5 0 5 Distance from the vein center (mm) 97.5

98 98.5 99 99.5 100

Extent of reaction (%)

Figure 1. Extent of reaction as a function of distance from the serpentine vein center. The extent of reaction illustrated in this figure is a maximum; it is calculated based on X-ray microto- mography (see section 2.1 for the computational method).

ray microtomography is a maximum due to these two constraints on olivine grain

173

identification.

174

2.2 Electron microprobe analysis and X-ray maps

175

Point analyses and X-ray maps were acquired with a JEOL JXA-8530F Hyper-

176

Probe at the University of Lausanne. All the data were acquired at 15 keV. The X-ray

177

maps were acquired at 40 nA with a dwell time of 50 ms and a resolution of 8.5μm.

178

Spot analyses of olivine, serpentine/brucite mixture and magnetite were acquired in

179

the same region as the maps to produce internal standards for converting the count

180

number of the X-ray maps into composition in oxide mass and then in mole fraction.

181

A mineral phase was then attributed to each pixel of the map based on composition

182

criteria (Figure 2). Point analyses reveal that serpentine and brucite are intermixed at

183

a scale of less than 1μm. The proportion of serpentine and brucite at each pixel identi-

184

fied as a serpentine/brucite mixture was determined assuming (M g, F e)3Si2O5(OH)4

185

and (M g, F e) (OH)2 as structural formulas for serpentine and brucite, respectively:

186

XBr=nMg+nF e32nSi

nMg+nF e−nSi (1)

wherenMg,nF eandnSidenote the molar content of Mg, Fe and Si, respectively,

187

andXBr is the mole fraction of brucite.

188

The Mg number of olivine (XMgOl) is determined by averaging the composition of

189

the pixels attributed to olivine. The Mg numbers of serpentine and brucite (XMgS and

190

XMgB, respectively) are determined by fitting for the serpentine/brucite mixture, the

191

Mg number (XMg) as a function ofXBr at each pixel with the following relationship:

192

XMg = XBr

32XBrXMgB+3 (1−XBr)

32XBr XMgS (2)

Finally, the Mg numbers calculated for each phase are used to determine their mo-

193

lar volume with a linear combination of the end-member molar volumes taking the fol-

194

(7)

0.76 0.8 0.84 0.88 0.92 0.96

Mg/(Mg+Fe)

A B

Figure 2. Compositional maps of sample BA4A-81-1-1-17 obtained from X-ray maps ac- quired with the microprobe. A: Mg/(Mg+Fe) ratio. The white arrow points towards the main serpentine vein. The black arrows point towards the core of a mesh preserved after complete serpentinisation. B: phases. The serpentine/brucite mixture is displayed in red, olivine in orange, spinel in blue and magnetite in green. The scale bar is 1 mm.

(8)

Figure 3. Standard deviation ofXBr as a function of the logarithm of the sampling size (sl) and the distance from the main vein (l).

lowing values: 43.79, 46.39, 24.63, 26.43, 108.5, and 115cm3.mol−1for forsterite, fay-

195

alite, Mg-brucite, Fe-brucite, Mg-serpentine and Fe2+-serpentine, respectively (Chichagov

196

et al., 2001).

197

To track changes in composition during reaction (export or import of aqueous

198

species), we calculate the mean composition of the serpentine/brucite mixture (XmBr)

199

by weighting the mole contents in brucite and serpentine at each pixel with their mo-

200

lar volume. Averages are calculated for all the pixels identified as serpentine/brucite

201

mixtures in the maps. We also estimateXmBr with a smaller sampling size to deter-

202

mine the scale at which compositional changes occur (scale of homogenization). The

203

side length of the square sampling box (sl) varies from the pixel size (8.5μm) to the

204

map width (5.6mm; Figure 3). We generate 3,000 randomly located sampling boxes

205

for eachsl. We then sort these boxes as a function of the distance to the main vein

206

network (l) to also determine the variation of the standard deviation ofXmBr with

207

sland l.

208

2.3 Estimate of volume change during reaction

209

In the sample studied here, primary olivine contains thin mineral platelets of

210

magnetite and clinopyroxene. Such inclusions are generally formed by exsolution at

211

high temperature during olivine cooling and oxidation (thus prior to serpentinisation;

212

(9)

Olivine Serp/Brc

Olivine Olivine

Serp/Brc

A B C ǻ9

9 = 50 % ǻ9

9 = 10 %

Figure 4. Sketchs illustrating the effect on inert markers displacement of solid volume change during the formation of a serpentine/brucite mixture (Serp/Brc). A: initial distribution of straight and parallel inert markers in olivine. B and C: predicted displacement of the inert markers for changes in volume of 10 % and 50 %, respectively. The three main effects of volume change during reaction for inert markers are: bending at the olivine/reaction products contact (red arrows), rotation (white arrows) and increase of the spacing (blue arrows).

(Moseley, 1984; A. Putnis, 1979)). They display a topotaxial intergrowth with olivine

213

leading to parallel oriented platelets along a preferred olivine crystallographic orien-

214

tation (Zhang et al., 1999; Ashworth, 2000). Baronnet and Boudier (pers. comm.)

215

first proposed use of such non-reactive and oriented markers to quantify solid volume

216

change during serpentinisation (Figure 4). Their calculation is based on the measure-

217

ment of the angle change between the olivine surface and an initially straight marker.

218

Such markers are also expected to rotate and to space out as a result of volume change

219

during reaction (Figure 4). Here we extend the approach of Baronnet and Boudier

220

(pers. comm.) by considering these latter effects to estimate solid volume change by

221

measuring the displacement of several platelets of magnetite and clinopyroxene.

222

First, we determine the deformation and composition of these platelets in three

223

dimensions with focused ion beam tomography at the CIME-EPFL laboratory (Lau-

224

sanne). We use a Zeiss Crossbeam 540 apparatus equipped with a field emission scan-

225

ning electron miscroscope. We erode layer by layer with an ion beam a 18x15x19μm

226

cuboid containing a deformed platelet and located at the surface of an olivine grain.

227

After each milling of a 10 nm thick layer, a back-scattered electron image is acquired

228

at an acceleration voltage of 1.7 kV with a current of 2.5 nA, a dwell time of 15 μs

229

per pixel and a pixel size of 10x10 nm. The stack of images is then aligned with the

230

FIJI image processing software by running the MultiStackReg plug-in (Th´evenaz et

231

al., 1998). We segment the images based on pixel intensity with a home-made Matlab

232

code by first applying a median filter and then using dilatation/erosion techniques.

233

As the inclusions are parallel in each olivine grain, we can compare orientations in

234

olivine and in fully serpentinised domains around each olivine. We therefore estimate

235

solid volume change in two dimensions by integrating the displacement of several

236

inclusions. We acquire optical photomicrographs of primary olivine surrounded by

237

serpentine + brucite. We select olivine grains with surface(s) oblique to the inclusion

238

orientations and with inclusions both inside the grain and in the secondary phases

239

(Figure 5 A). The olivine rim is mapped in two dimensions with a set of segments.

240

We then compute regularly spaced lines parallel to these segments (spacing dini =d

241

between the lines) to generate a mesh (Figure 5 B). The inclusions outside the olivine

242

(10)

A B

C D

Ol

Serp/Brc Inclusion

Figure 5. Steps for determining solid volume change during serpentinisation. We measure the misorientation of inclusions contained in olivine (Ol) initially and in the serpentine/brucite mixture (Serp/Brc) after reaction. A: an olivine grain surrounded by serpentinisation reaction products is first selected. The grain surface must be oblique to the inclusions. B: the inclu- sions are mapped (red lines outside the olivine grain and blue lines inside the olivine grain) and projected on a 2-D mesh generated from the olivine surface. C: the inclusions are deformed by modifying the mesh spacing in the direction normal to the olivine grain surface. D: the best model (green lines) is selected from a set of models with different mesh spacing by minimizing the angle between the needles inside and outside the olivine grain using the least-squares method.

Comparing initial mesh spacing to the spacing in the best model provides an estimate of solid volume change during reaction (see text for details).

(11)

grains are projected on the mesh which can be deformed by modifying d. The least-

243

squares method is finally applied to determine the value of d, dbest, for which the

244

orientation of the needles outside the olivine grain best fits the orientation of the

245

needles inside the olivine grain (Figure 5D). The surface area of the mesh at dbest

246

(Sbest) can finally be compared to the surface area of the mesh before deformation

247

(Sini) to estimate the change in surface area during reaction (ΔSS = SiniSSbest

best ). Thin

248

sections obtained from perpendicular sections of the sample display the same mesh

249

textures with reaction zones of approximately uniform size around the grains. We

250

therefore assume homogeneous volume change for 3D calculations in the following.

251

For spherical grains, ΔSS does not depend on the orientation of the 2D thin section

252

used for measurement. The change in surface area can be converted in a change in

253

solid volume (ΔVV) with the following equation:

254

ΔV V =

ΔS S

1(1−ξ)2/3

+ 1 3/2

1

ξ (3)

withξthe local extent of reaction. ξcan be estimated from 2D images as:

255

ξ= 1 Sol

Sbest 3/2

(4)

withSol the surface area of the olivine grain. For spherical grains, Equation 4

256

only provides the actual extent of reaction if the grains are cut through their centers.

257

It overestimates ξ otherwise. Therefore, ΔVV determined with ξ from equation 4 is

258

a maximum. The limit of equation 3 as ξ tends towards zero is ΔVV = ΔSS, which

259

provides a minimum estimate for the change in volume.

260

2.4 Laser-ablation inductively coupled plasma mass spectrometry (LA-

261

ICPMS)

262

The chemical data are acquired with a sector-field Element XR inductively cou-

263

pled plasma mass spectrometer (ICP-MS) coupled with an Atlex 193 nm ArF excimer

264

laser housed inside an Australian Scientific Instruments (ASI) RESOlution system (in-

265

cluding an S155 dual volume sample cell) at the University of Lausanne. We use two

266

strategies to acquire the data: spot mode and continuous scanning. Both datasets are

267

acquired along the direction perpendicular to the main vein network. For each ele-

268

ment, one isotope is measured in a peak-hopping mode. The on-sample laser energy

269

densities are approximately 10 J.cm−2 with repetition rates at 10 Hz. The laser beam

270

shape of the spot mode is rectangular with a size of 100 x 25 μm. For continuous

271

analyses, the laser beam is a disk with a diameter of 100μmand the scanning rate is

272

10μm.s−1.

273

To reduce the results, glass standard BCR-2G was analyzed along with the sam-

274

ples to be used as an external calibration standard, while silica (SiO2 = 33 wt.%) is

275

used as an internal standard using LAMTRACE (Jackson, 2008) and Iolite (Version

276

3.11, The University of Melbourne) (Paton et al., 2011) software. An interval of 1.58

277

s is selected in the data reduction for continuous scanning. For smoothing and clarity,

278

we use a moving average on three points for plotting the results. The raw geochemical

279

data are given in the Supplementary Materials.

280

(12)

3 Results

281

3.1 Mineralogical distribution

282

The sample is cross-cut by a 5mm wide anastomosing network of serpentine

283

veins (referred to as ”black veins” in the core description (Kelemen et al., n.d.)). De-

284

spite complete serpentinisation in the vicinity of these veins, the typical mesh texture

285

of serpentinised peridotites is preserved (Figure 2 A). Approximately 10μmwide ser-

286

pentine veins surround mesh cores composed of a concentric alternation of brucite-rich

287

and serpentine-rich layers. Raman spectroscopy reveals that the serpentine minerals

288

are both chrysotile (main peak at 3699 cm−1and shoulder at 3689 cm−1) and lizardite

289

(main peak at 3684 cm−1 and secondary peak at 3705 cm−1) in the vein network. In

290

the mesh cores, lizardite is intermixed on the sub-micrometer scale with brucite (main

291

peak at 3636 cm−1 and secondary peak at 3652 cm−1). This latter peak corresponds

292

to the peak expected for pure brucite (M g(OH)2; (Speziale et al., 2005)). The shifting

293

towards lower frequencies of the main brucite peak suggests the presence ofF e(OH)2

294

(Speziale et al., 2005). Magnetite is only found as sub-micron sized grains dissem-

295

inated in serpentine and brucite. It mainly occurs in the wider serpentine veins of

296

the anastomosing network. Up to2 mm wide chromites are found throughout the

297

samples. They can react at their margin with the development of a 2μmwide reaction

298

zone mainly composed of magnetite. We also observe minor 50 μm wide clinopy-

299

roxenes in the rock matrix. They do not display evidence for reaction in agreement

300

with thermodynamic calculations (Klein et al., 2009).

301

3mmaway from the anastomosing vein network, olivine grains are found in the

302

mesh core (Figure 2). The transition from olivine-free to olivine-bearing serpentinised

303

peridotites occurs at approximately 4 mm from the main serpentine vein (Figure 1).

304

The extent of reaction is generally high with a minimum estimated at 97 % with X-ray

305

microtomography and at 80 % with X-ray maps acquired with the microprobe.

306

Olivine grains commonly contain 1μm wide, parallel inclusions (Figure 6).

307

We use these platelets to estimate solid volume change during serpentinisation in

308

section 3.4. Energy dispersive spectroscopy reveals that the inclusions are intergrowths

309

of clinopyroxene and magnetite, containing minor concentrations of Cr. FIB-SEM

310

nanotomography shows that the platelets have a symplectic texture (Figure 7 A). Its

311

composition is similar when it is included in the olivine or in the serpentine/brucite

312

mixture and thus the composition of both clinopyroxene and magnetite appear to

313

be largely unaffected by serpentinisation. The only effect of reaction appears to be

314

the generation of discontinuities in the platelet separating 1 μm long segments of

315

clinopyroxene. The presence of traces of magnetite in the discontinuities as well as

316

their orientation in parallel to other magnetite grains suggest that the discontinuities

317

were formerly filled with magnetite which reacted during serpentinisation.

318

3.2 Mineralogical composition

319

Point EPMA analyses and X-ray maps indicate homogeneous iron distribution

320

in olivine with a Mg number (MgMg+F e mole fraction) of 0.895±0.011. NiO content in

321

olivine ranges from 0.25 to 0.33 wt.%.

322

For the brucite/serpentine mixture, XMg varies as a function of XBr along a

323

mixing line defined by equation 2 (Figure 8). The fit of the data leads to XMgS =

324

0.956±0.022 andXMgB = 0.724±0.022. XBr averaged over the entire mapped area

325

is 0.515±0.002. The mean value ofXMg in the serpentine/brucite mixture is 0.896.

326

Figure 3 gives the standard deviation ofXBr as a function of the sampling size (sl)

327

and the distance to the main serpentine vein (l). It thus provides an estimate of the

328

scale and extent of mass transfer in the sample. The standard deviation decreases

329

withsl and is high for sl below approximately 500 μm. Below this sl, the standard

330

(13)

20 μm

Ol Serp/Brc

Mag-Cpx intergrowth Mag

Figure 6. Back-scattered electron image of an olivine grain (Ol) reacted at its border to form a serpentine/brucite mixture (Serp/Brc). Olivine contains parallel inclusions of magnetite(Mag)- clinopyroxene(Cpx) intergrowths which are displaced in the serpentine/brucite mixture.

deviation highly depends onl and is 2 to 3 times higher in the first 2 mm near the

331

main serpentine vein than further away where olivine also occurs in the matrix (Figure

332

3).

333

3.3 Chemical composition

334

Compositional profiles, measured from both X-ray maps and LA-ICPMS, display

335

constant contents of FeO, MgO and SiO2 at a scale larger than several hundreds of

336

micrometers which is also the scale at which constant brucite contents are measured.

337

Figure 2 A reveals variations of more than 15 % in the Mg/(Mg+Fe) ratio at a smaller

338

scale and near the main serpentine vein. This is associated with brucite and serpentine

339

segregation in the mesh cores.

340

Spatial variations in Al2O3 and Cr2O3 are mainly associated with the primary

341

distribution of chromite in the sample. From a trace element point of view, there are

342

three distinct correlations between the abundances and the distances from the main

343

vein. Concentrations of fluid-mobile elements such as Na, Ca, Rb, Sr and Ba gradually

344

decrease with increasing perpendicular distance away from the main vein. Some spikes

345

in Ca and Na profiles may be due to the preservation of fine clinopyroxene grains or

346

inclusions (Figure 9).

347

Some elements display constant contents, such as B, Co, Ni, Cu, Zn and Yb

348

(Figure 10). Most of them are transition metal or heavy rare earth elements (HREEs).

349

The light-REEs, Nb, Pb and U are very low and their variation is difficult to

350

quantify (for example, the contents of Ce mostly cluster around 1 ppb, ignoring the

351

spikes possibly related to clinopyroxene). U is generally less than 1 ppb and provides

352

a possible line of evidence for the absence of carbonate formation. Finally, Al, Sc, Ti,

353

V and Sr show periodic increases at a length scale of 5 mm (Figure 11). The bulges

354

formed by these increases are rather symmetric with a flat plateau at the highest

355

concentration surrounded by sharp flanks where the concentration decreases.

356

(14)

Cpx

Ol Serp/Brc

A B

Figure 7. FIB-SEM nanotomography of a platelet which has partly experienced serpentinisa- tion. A: result of segmentation. The contact between olivine (Ol) at the bottom and the serpen- tine/brucite mixture (Serp/Brc) at the top is displayed with the yellow surface. The platelet is composed of clinopyroxene (Cpx; green) and magnetite (Mag; gray). It is displaced along cracks at some magnetite/clinopyroxene contacts in the serpentine/brucite mixture (red arrow). It is also discontinuous at its top (white arrows). B: angle between a reference orientation and the normal to the platelet surface. The reference orientation is chosen as the normal to the platelet when it is still included in the olivine grain (bottom part).

(15)

0 0.2 0.4 0.6 0.8 1 Brucite molar fraction (X

Br) 0.75

0.8 0.85 0.9 0.95

Mg number (X Mg)

2.4 2.6 2.8 3 3.2 3.4 3.6

Figure 8. Mg/(Mg+Fe) ratio (Mg number) of the serpentine/brucite mixture as a function of the brucite molar fraction (XBr) calculated with equation 1. Data are obtained at each pixel identified as a serpentine/brucite mixture. The density of data points is contoured (arbitrary units). The black line corresponds to the best fit of the data obtained with equation 2. 95 % and 68 % confidence bands are displayed with red and pink plain lines, respectively. We used the 68

% confidence bands to estimate the uncertainty onXMgS andXMgB.

(16)

Ol relics appear

Fine Cpx Fine Cpx

0 1000 2000 3000 4000 5000 6000 7000 8000 9000

Sr ppm

89 7 6 5 4 3

'LVWDQFHIURP9HLQȝP

Rb ppb 100

101

10-1

Ca 103 ppm

1.0 2.0 3.0

0.7

103 104 Ca ppm

Na ppm

30 40 50

20

10

Figure 9. Concentration of water soluble elements (Na, Ca, Rb and Sr) as a function of the distance to the main vein network. The profiles are acquired perpendicular to the main vein

(17)

Ol relics appear 1

10 50

B ppm

101 102 103

Co ppmNi 103 ppm

1 2 3 4

0 1000 2000 3000 4000 5000 6000 7000 8000 9000

Yb ppm

'LVWDQFHIURP9HLQȝP 10-1

10-2

Figure 10. Concentration of B, Co, Ni and Cr as a function of the distance to the main vein network. The profiles are acquired perpendicular to the main vein network. The first appearance

(18)

Original Ol grain Original Ol grain

10-1

V ppm

100

Ti ppm

30 25 20 15

10

0 1000 2000 3000 4000 5000 6000 7000 8000 9000

101 102

Cr ppm

'LVWDQFHIURP9HLQȝP 100

Al ppm

102

101

Figure 11. Concentration of Al, Ti, V and Sc as a function of the distance from the main vein network. The profiles are acquired perpendicular to the main vein network. The first ap-

(19)

Figure 12. Change in solid volume (ΔVV) measured in sample BA4A-81-1-1-17. Minimum (blue bars) and maximum (orange bars) values of ΔVV are estimated for each photomicrograph.

The procedure for calculating these minimum and maximum bonds is described in the Method section. The mean (bold lines)±the standard deviation (thin lines) are displayed with dashed and plain lines for the minimum and maximum values of ΔVV, respectively. The bold dotted line corresponds to the change in volume calculated when considering no mass transfer during reaction.

3.4 Solid volume change

357

We used the misorientation of clinopyroxene/magnetite intergrowths initially in-

358

cluded in olivine to quantify volume change during serpentinisation. The platelet

359

analysed with FIB-SEM tomography is planar when it is included in olivine with less

360

than 5 of misorientation (Figure 7 B). It progressively bends when it is included in

361

the serpentine/brucite mixture (Figure 7 A and B) with misorientation values of more

362

than 15 the further away from the olivine surface. Bending preferentially occurs

363

along small cracks that are mostly located at clinopyroxene/magnetite contacts. The

364

acute angle between the platelet and the olivine surface increases with this bending.

365

This indicates a positive change in solid volume during reaction. Systematic differ-

366

ences in the orientation of ensembles of inclusions, between those hosted in olivine and

367

those hosted in surrounding zones of serpentine + brucite replacing olivine, are also

368

observed in scanning electron images (Figure 6).

369

To extract quantitative estimates of solid volume change during reaction, we use

370

the procedure described in section 2.3. We analyse 64 optical photomicrographs of

371

serpentinised olivine. The results of inclusion reorientation are summarized in Figure

372

12 and provided for each photomicrograph in the Supplementary materials. We mea-

373

sure positive ΔVV from 8 to 168 %. ΔVV minimum and maximum bonds have mean

374

values of 59 ± 29 and 74 ± 36 %, respectively (see section 2.3 for the definition of

375

these bounds).

376

(20)

4 Discussion and conclusion

377

The sample studied here appears to be ideal for studying mass transfer and

378

volume change during reaction. It comes from depth (215 m) and has thus experienced

379

relatively limited surface alteration; for example, we do not observe carbonates or

380

evidence for carbonation in U, Sr or Ca datasets consistent with the results of shipboard

381

core description (Kelemen et al., n.d.). Moreover, the NiO content of the remaining

382

olivine is below 0.32 wt. % indicating that the protolith is probably a dunite (Suhr,

383

1999).

384

Serpentinisation requires transport of 150 g of water per kilogram into an

385

initially impermeable peridotite. The sample is crossed by an anastomosing vein per-

386

pendicular to which a gradient in extent of reaction is measured (Figures 2 and 1). This

387

suggests that water migrated from the main serpentine vein into the peridotite matrix.

388

The observed mesh texture on the wall of the main vein indicates that a fracture net-

389

work with a typical spacing of 100μmpromoted fluid transport inside the peridotite.

390

Such an organised crack network may have formed by reaction-induced fracturing

391

(Jamtveit et al., 2009; Kelemen & Hirth, 2012; Pl¨umper et al., 2012; Malvoisin et al.,

392

2017) and/or thermal cracking (Demartin et al., 2004; Boudier et al., 2005). The origin

393

of the main serpentine vein is more difficult to constrain since tectonic processes could

394

also promote fracturing (Roum´ejon & Cannat, 2014). The fluid transport direction

395

is confirmed with compositional profiles in trace elements. A decrease in fluid-mobile

396

elements (Na, Ca, Sr, Rb and Ba) is indeed observed from and perpendicular to the

397

main vein. The history of the Oman ophiolite is complex with four main events dur-

398

ing which serpentinisation probably occurred: hydrothermal circulation at mid-ocean

399

ridges, interaction with metamorphic fluids during obduction (Searle & Malpas, 1980),

400

Cretaceous weathering of the ophiolite, followed by a marine transgression (de Obeso

401

& Kelemen, 2018), and modern ongoing low temperature serpentinisation through in-

402

teraction with rainwater (Neal & Stanger, 1983; Kelemen & Matter, 2008; Kelemen et

403

al., 2011). Fluids collected at surface seeps or in wells in the Oman ophiolite generally

404

display high Na, Cl, Sr and Ca concentrations (Chavagnac et al., 2013; Paukert et al.,

405

2012; Canovas et al., 2017; Paukert Vankeuren et al., 2019), the same as the elements

406

that display a progressive decrease from the main vein in the sample studied here.

407

Modern water sources and seawater both contain all these components at relatively

408

high concentration (Millero et al., 2008; Paukert Vankeuren et al., 2019). The time of

409

serpentinisation can thus not be determined with the data presented here. The enrich-

410

ment in fluid-mobile components in the Oman ophiolite fluids could be secondary and

411

acquired through dissolution of fluid-mobile elements previously trapped in gabbro or

412

in peridotites serpentinised on the seafloor (Miller et al., 2016). Meanwhile, the Sr

413

isotope ratios in peridotite-hosted carbonates, water, and leachates are substantially

414

more radiogenic than Cretaceous seawater (Kelemen et al., 2011). Several sources may

415

thus contribute to the modern composition of the fluids collected in Oman.

416

The standard deviation of the Mg/(Mg+Fe) ratio in the serpentine/brucite mix-

417

ture is three times higher for a sampling size below 100-500 μm than above (Figure

418

3). This indicates that major elements are redistributed at a scale below 100-500μm

419

corresponding to the mesh scale. This redistribution increases near the main vein

420

(Figure 3) indicating a larger scale segregation of brucite and serpentine. There is

421

thus mass transport probably associated with further fluid-rock interaction at a scale

422

of less than several hundreds of micrometers. This local mass transfer is probably the

423

incipient stage of the mass transport responsible for the change in chemical compo-

424

sition observed in serpentinised peridotites with a decrease of the M gO/SiO2 ratio

425

during more extensive alteration (Snow & Dick, 1995; Niu, 2004; Monnier et al., 2006;

426

de Obeso & Kelemen, 2018).

427

We do not measure spatial changes in major element composition at the cen-

428

timeter scale in microprobe and LA-ICPMS data. The Mg/(Mg+Fe) ratio in the

429

(21)

serpentine/brucite mixture is 0.896, similar to the one measured in the remaining

430

olivine grains (0.895±0.011). Brucite and serpentine display an equimolar distribu-

431

tion in the serpentine/brucite mixture (XBr = 0.515±0.002) and magnetite is rare

432

in the investigated area (Figure 2 B). This suggests that the initial (Mg+Fe)/Si ra-

433

tios of olivine are conserved during reaction. Excluding trace elements and at a scale

434

higher than several hundreds of micrometers, serpentinisation in the studied sample

435

corresponds to the isochemical hydration of olivine:

436

(M g0.90F e0.10)2SiO4+3H2O= (M g0.96F e0.04)3Si2O5(OH)4+(M g0.72F e0.28) (OH)2 (5)

where the compositions of serpentine and brucite are retrieved from microprobe

437

data with equation 2. The Mg numbers of brucite and serpentine can theoretically pro-

438

vide constraints on temperature and water to rock ratio. However, the uncertainties

439

on the thermodynamic parameters of Fe-brucite and Fe-serpentine preclude precise

440

quantitative estimates. The XMgB measured here is at the low end of the range for

441

thermodynamic predictions (Klein et al., 2009, 2013). The thermodynamic calcula-

442

tions indicate that more iron is incorporated in brucite at low temperature (T<100

443

C) and low water to rock ratio (< 0.5). The scarcity in magnetite in the sample is

444

also compatible with such conditions.

445

Al, Ti, V, Sc and Cr concentrations periodically vary at a typical length scale

446

of several millimetres without relationship to the extent of serpentinisation (Figure

447

11). Measured compositional profiles recall the diffusion profiles for the same ele-

448

ments measured in unaltered olivine-bearing xenoliths at a length scale typical of the

449

olivine crystal size in unaltered peridotites (Tollan et al., 2015). Therefore, we inter-

450

pret the measured composition as reflecting the primary olivine composition acquired

451

through diffusion or growth at high temperature. Preserving olivine primary composi-

452

tion requires that Al, Ti, V, Sc and Cr are immobile during the whole serpentinisation

453

process.

454

Our study of major and trace elements behaviour during serpentinisation indi-

455

cates very limited mass transfer except for elements associated with the fluid (Na, Ca,

456

Sr, Rb and Ba). In contrast, several studies have reported a decrease in MgO/SiO2in

457

serpentinised peridotites, compared to their mantle peridotite protoliths (Snow & Dick,

458

1995; Niu, 2004; Monnier et al., 2006; Malvoisin, 2015; de Obeso & Kelemen, 2018).

459

This decrease is either related to Mg loss during alteration on the seafloor (Snow &

460

Dick, 1995) or to Si metasomatism (Paulick et al., 2006; de Obeso & Kelemen, 2018).

461

Thermodynamic calculations predict a significant decrease in the M gO/SiO2 ratio

462

only when consideringSiO2gain. Mg loss is not predicted at equilibrium even at high

463

water to rock ratio (105) (Malvoisin, 2015). However, high-resolution tomography in-

464

dicates brucite dissolution and porosity formation in abyssal serpentinised peridotites

465

exposed on the seafloor (J¨ons et al., 2017). Mg loss would then be associated with

466

kinetic effects limiting the precipitation of Mg-bearing phases (e.g. carbonates) during

467

fluid flow. Brucite is not commonly observed in the samples dredged on the seafloor

468

and studying its fate in the cores drilled in Oman can thus provide insights into the

469

decrease in theM gO/SiO2 ratio measured in abyssal peridotite. As brucite contains

470

approximately 30 % iron, its reaction is expected to produce iron-bearing phases such

471

as magnetite while magnesium is either leached from the system with the fluid (Snow

472

& Dick, 1995) or combined with dissolved silica to form serpentine (Beard et al., 2009).

473

In our study, we observe tiny magnetite grains (<1μm) that could be associated with

474

incipient brucite alteration at low temperature (Figure 2). Magnetite production is

475

also expected to produce hydrogen according to experiments on samples from Oman

476

(Miller et al., 2017).

477

Références

Documents relatifs

Onset and endpoints (dots) of modeled abrupt transitions (oblique lines) together with associated uncertainty intervals (horizontal shaded lines) found by the ramp- fitting analysis

(e) Interpretative cross-section reconstructed in Cretaceous, to illustrate the SE-ward progradation of the platform margin from late Jurassic-Berriasian (Échaillon quarry,

The four ma- jor aims of the SCOPSCO project are to (i) obtain more in- formation on the age and origin of Lake Ohrid, (ii) unravel the regional seismotectonic history including

Remarks: This species differs from Cerastus praeinsularis Neubert and Van Damme, 2012 by its lesser dimensions, the distinctly lower last whorl, the strongly convex spire whorls and

In these loops the data generated either by the mud logging or the logging while drilling are processed to gener- ate information that can be used directly by the drilling engi-

(a) In a basin of the Windward Passage along the OSFZ offshore the northern Haitian peninsula, (b) in the Navassa Basin of the Jamaica Passage along the EPGFZ offshore the

(b) highly thinned continental or oceanic crust overlying ser- pentinized mantle; and (c) exhumed lower crust or a mixture of lower crustal and upper mantle material.. The

This ‘lightweight’ ice-core drilling project, staffed by a maximum of eight people at the drill site, successfully recovered an ice core to 948 m depth from a remote Antarctic site