Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal circulation

Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal circulation

Journal Pre-proof Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal ci...

2MB Sizes 0 Downloads 45 Views

Journal Pre-proof Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal circulation. Thalia Avilés Esquivel, Carlos Flores, Valeria Reyes Ortega, Steven Constable, Enrique Gómez-Treviño, Antonio González-Fernández PII:

S0895-9811(20)30014-6

DOI:

https://doi.org/10.1016/j.jsames.2020.102501

Reference:

SAMES 102501

To appear in:

Journal of South American Earth Sciences

Received Date: 8 July 2019 Revised Date:

10 January 2020

Accepted Date: 12 January 2020

Please cite this article as: Esquivel, T.A., Flores, C., Ortega, V.R., Constable, S., Gómez-Treviño, E., González-Fernández, A., Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal circulation., Journal of South American Earth Sciences, https://doi.org/10.1016/j.jsames.2020.102501. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.

December 14th, 2019 To whom correspond, I attach the author’s statements, where I describe the contribution of each one of the authors to this publication.

Authors Thalia Anaid Avilés Esquivel (Corresponding author) Carlos Flores Luna

Formal Analysis, visualization, validation, writing the original draft Conceptualization, Investigation, writing the original draft

Valeria Reyes Ortega

Formal analysis, validation, visualization

Steven Constable

Methodology, software, resources, supervision

Enrique Gómez Treviño

Methodology, software

Antonio González Fernández

Proyect administration, found resources, funding acquisition.

Best regards,

Thalia Anaid Avilés Esquivel

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

Magnetotelluric exploration of the Wagner Basin, Gulf of California, Mexico: Evidence for an axial magma chamber and hydrothermal circulation. Thalia Avilés-Esquivel a,b,*, Carlos Flores a,b, Valeria Reyes-Ortega a,b,d , Steven Constable c, Enrique Gómez-Treviño a,b and Antonio González-Fernández a,b a

División de Ciencias de la Tierra, Centro de Investigación Científica y de Educación Superior de Ensenada (CICESE), Carretera Ensenada-Tijuana 3918, Zona Playitas, Ensenada, Baja California 22860, México. b Centro Mexicano de Innovación en Energía Geotérmica (CeMIEGeo), Carretera Ensenada-Tijuana 3918, Zona Playitas, Ensenada, Baja California 22860, México. c Institute of Geophysics and Planetary Physics, Scripps Institution of Oceanography, University of California San Diego, La Jolla, 92093, USA. d Now at c * Corresponding author: [email protected] Keywords: Marine Magnetotellurics, electrical resistivity – heat flow correlation Abstract

24

Different geological and geophysical disciplines have suggested the Wagner Basin as a

25

promising location of geothermal resources. As a reconnaissance tool Magnetotelluric

26

(MT) data were measured at 10 sites along a profile over this basin in the northern Gulf

27

of California. The subsurface distribution of electric resistivity was estimated with two-

28

dimensional smooth inversion using the apparent resistivities and phases from both field

29

polarizations. We found a deep conductor underneath the center of the basin,

30

interpreted as a region of partial melt feeding the accretion zone of this incipient oceanic

31

crust. From six MT sites with close-by heat flow measurements, we also found a

32

positive correlation between the interpreted shallow (first 100 m) low resistivities and

33

high heat flows. This shows that the MT method can be used to map submarine heat

34

sources.

35 36

Introduction

37

The Wagner Basin is the northernmost rifting basin in the Gulf of California.

38

Several facts suggest this basin as a potential source of geothermal resources. First, its

39

tectonic similarity with the Mexicali-Imperial valley to the north, which has important

40

geothermal fields such as Cerro Prieto, Salton Sea, Brawley, East Mesa, and Heber.

41

The combined installed capacity of these fields in 2015 was 1.5 GWe (Boyd et al., 2015;

42

Gutiérrez-Negrín et al., 2015). Second, the high temperatures and heat flow

43

measurements reported in this area since the earliest 70’s (Henyey and Bischoff, 1973;

44

Grijalva, 1986; Prol-Ledesma et al., 2013; Neumann et al., 2017). The measured heat

45

flows in the Wagner and neighboring Consag basins are up to 15 times the average

46

value for oceanic crust. Third, the intense submarine gas discharges reported by Canet

47

et al. (2010) and the geochemical signature of high-temperature pore water measured

48

in sediment cores by Batista Cruz et al. (2019). The mapped seismic events (González-

49

Escobar et al., 2010) and the active distributed deformation (Persaud et al., 2003)

50

suggest the presence of active tectonism. Finally, the incipient formation of new crust is

51

suggested by the possible presence of intrusive bodies in a seismic reflection profile

52

over the Consag Basin (González-Escobar et al., 2014) and in the gravity anomaly in a

53

profile over the Wagner Basin (Pérez, 1982).

54 55

Regarding the geologic setting, the Gulf of California – San Andreas fault system

56

defines the boundary between the Pacific and North American plates (Atwater, 1970;

57

Lonsdale, 1989). The Wagner Basin is the northernmost rifting basin in the Gulf of

58

California. The Consag, Upper and Lower Delfín, and Guaymas are among other basins

59

to the south. Important differences exist between the northern and southern Gulf basins.

60

Well developed spreading centers occur at the mouth of the Gulf, with clearly

61

observable magnetic anomaly stripes, evidence of new oceanic crust being created

62

since about 4 Ma at a full spreading rate of 6 cm/year (Larson et al., 1968). In contrast,

63

in the northern part no magnetic lineaments are observed (Klitgord et al., 1974;

64

Sanchez et al., 1991) and thick sedimentary deposits, supplied mainly by the Colorado

65

River, fill these basins (Dorsey, 2010). The lack of magnetic anomalies in the northern

66

basins is the main evidence for the absence of new oceanic crust. However, the region

67

of the northern basins is considered a continent-ocean transition zone, where diffuse

68

continental deformation exists (Nagy and Stock, 2000; Persaud et al., 2013).

69

70

In this work, we present the results of a Magnetotelluric (MT) reconnaissance

71

survey carried out in the southern portion of the Wagner Basin (Figure 1) with the

72

purpose of assessing its geothermal potential. This project is part of a national effort to

73

increase the use of renewable energy sources (geothermal, wind, solar, etc.) to diminish

74

the hydrocarbon dependence (Avilés Esquivel, 2016). Technologic and environmental

75

obstacles should be overcome for future sustainable exploitation of the submarine

76

geothermal resource in this environmental protected zone.

77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93

94

95 96

Figure 1: Regional map of the northern Gulf of Baja California. Lines denote the normal

97

faults inferred from seismic reflection data from PEMEX (Petróleos Mexicanos) (Martín-

98

Barajas et al., 2013). CF=Consag Fault, WF=Wagner Fault, CPF=Cerro Prieto Fault,

99

CPSF=Cerro Prieto South Fault, ADF=Adair Fault, AMF=Amado Fault, WB=Wagner

100

Basin, CB=Consag Basin. Black dots indicate the location of the MT sites. The

101

bathymetry is from G. Díaz Méndez (pers. comm.). The boundaries between the Pacific

102

(PP) and North American (NAP) plates, as well as their direction of displacement are

103

shown in the inset.

104 105

106

The use of electromagnetic (EM) methods to study the marine subsurface has

107

shown an important increase in the last years (Key, 2012). These works have either

108

academic or resource exploration goals, mainly for the characterization of offshore

109

hydrocarbon reservoirs, but also for studying the oceanic lithosphere (e.g., Evans et al.,

110

1991; Constable et al., 1997; Heinson et al., 2000; Baba et al., 2006a; Key et al., 2013).

111

The MT method uses low-frequency EM waves originated by sun-ionosphere

112

interactions and far-away lightning sources. These waves impinge upon the sea

113

surface, diffusing into the seawater and seabed. Measured electric and magnetic fields

114

on the seafloor are then used to estimate the frequency-dependent impedances, which

115

are inversely modeled to estimate the subsurface electrical resistivity. The presence of

116

a thick sedimentary layer in the Gulf of California dissipates the signal from seismic and

117

gravity sources, preventing to obtain deeper information about the characteristics of the

118

underlying high velocity and denser structure. Using MT marine studies, we will

119

penetrate below the thick sedimentary layer and describe for the first time the electrical

120

structures related to a possible intrusive magma body.

121 122

The relationships between high temperatures and electrical resistivity and their

123

implication with marine MT studies have not been studied. So we explore if MT marine

124

models are sensitive to detect changes in the shallow resistivity associated with high

125

heat flows. The previous results may allow us to have a better understanding of the

126

electrical resistivity influence on the high heat flows and their relationship with the

127

intrusive body. This approach might initiate a wider use of MT marine methods as an

128

instrument for assessing geothermal potential in this area.

129 130

Magnetotelluric data

131

We collected 10 marine magnetotelluric (MT) stations for 18 days in May 2015

132

over the Wagner basin (Figure 1). The data were acquired using an MT system

133

developed at the Scripps Institution of Oceanography (Constable et al., 1998), obtaining

134

earth response functions in the period band from 0.12 to 6000 s. The time variations of

135

the two horizontal magnetic and electric fields were recorded. The Ocean Bottom

136

Electromagnetic (OBEM) instruments were deployed along a profile consisting of 10

137

stations distributed along 80 km. This profile has an azimuth of 65o, it is approximately

138

perpendicular to both coasts of the Gulf of California (Figure 1), it starts near San

139

Felipe, Baja California, and ends near the town of Puerto Peñasco, Sonora. The

140

separation of the different sites is not uniform, varying from 7 to 15 km. The compasses

141

attached to each OBEM did not work properly. The orientations of the receivers at the

142

seafloor were estimated with cross-correlations of the magnetic field measured at the

143

Pinon Flat Observatory, southern California, with the rotated magnetic fields recorded at

144

the receivers. This observatory is about 350 km north from the MT profiles. The data of

145

one of the electric field components at Site 8 was lost before the OBEM recovery.

146 147

The impedance tensors were calculated using a robust, multi-station transfer-

148

function estimation routine (Egbert, 1997). Figure 2 shows the impedance polar

149

diagrams (Reddy et al., 1977) at the MT sites for some selected periods. These

150

diagrams plot the normalized absolute values of the impedances Z x y and Z x x for a

151

360o rotation angle. Figure 3 helps to discuss the main features of the data and displays

152

the observed apparent resistivities and phases for the transverse electric (TE) and

153

transverse magnetic (TM) modes obtained by rotating the impedances to an angle

154

coincident with the azimuth of this profile (65o).

155 156 157 158 159 160 161 162

163

164 165

Figure 2. Impedance polar diagrams of the MT sites for some selected periods (north

166

upwards). The absolute values of the impedances Zxy (blue lines) and Zxx (red lines) are

167

plotted as a function of the rotation angle.

168

The general behavior of the data can be divided into four-period ranges. From

169

0.13 to 0.5 s the Z x y diagrams are nearly circular (Figure 2), and the responses of the

170

two modes almost coincide (Figure 3), suggesting a 1D resistivity distribution associated

171

with the shallow portion of the basin sediments. The apparent resistivities are low,

172

varying from 0.3 to 0.9 Ωm, except at sites 9 and 10, where they are 3.1 and 2 Ωm.

173

From 0.7 to 11 s the polar diagrams suggest the presence of 3D subsurface structure;

174

however, the coherences between the fields were significantly lower in these periods.

175

We inferred that this band is contaminated by noise, probably the combination of low

176

intensity of the ionospheric signal and the effect of sea waves in this relatively shallow

177

sea. We discarded practically all the data contained within this period band. From 16 to

178

130 s, the strike directions are not homogeneous, but an NW-SE strike can be identified

179

as slightly dominant, and the apparent resistivities of the two modes are either

180

coincident or start to separate. These features suggest that the structure is mainly 2D.

181

Finally, for periods longer than 130 s the relative increase of Z x x , the shapes of Z x y ,

182

and the separation of the TE and TM modes, all suggest that these responses are the

183

product of a 3D structure.

184 185 186 187 188 189 190 191

192

193

Figure 3. Apparent resistivities ( Ω m) and phases (degrees) for the MT soundings. The

194

observed data are shown with blue dots (TE mode) and red circles (TM mode). The

195

bars indicate +/- one standard error. The responses calculated from the inverted model

196

displayed in Figure 4 are depicted as blue (TE mode) and red (TM mode) solid lines.

197

The Root Mean Squared (RMS) misfit error between calculated and measured

198

responses is 1.7

199 200

The data were inverted with the MARE2DEM program, a 2D code (Key, 2016)

201

that incorporates the smooth-structure approach (Constable et al., 1987) using an

202

unstructured finite-element method to calculate the forward problem (Key and Weiss,

203

2006; Key and Ovall, 2011). The bathymetry, measured with an echo sounder during

204

the research cruise (Díaz-Méndez, pers. comm.), was incorporated into the model. The

205

water depth along the profile varies from 31 to 208 m. The shallowest zone lies at the

206

eastern end of the profile, and the deepest region is located near sites 4 and 5, at the

207

center of the basin. Isotropic resistivity was considered in all inversions. An error floor of

208

10% in apparent resistivity and phases was considered.

209 210

Results

211

Figure 4b shows the inverted 2D model obtained after seven iterations using as

212

input data the apparent resistivities and phases for the TE and TM polarization modes.

213

The fit between observed and calculated responses (Figure 3) is good for periods less

214

than 400 s, but not so good for longer periods, with a value of 1.7 for the overall Root

215

Mean Squared (RMS) misfit error. Figure 4c displays the interpretation by Aragón

216

Arreola (2006) of a seismic reflection line almost coincident with our electromagnetic

217

profile.

218 219

Bouguer Anomaly (mGal)

220

221 222 223 224 225 226 227 228 229

Figure 4. a) Bouguer gravity anomaly (mGal) from Pérez (1982). b) 2D model estimated by the inversion of the MT data. The rms misfit error is 1.69. The main resistivity features are indicated: shallow conductor (SC), shallow resistor (SR), deep resistor (DR), deep central conductor (DCC), and deep eastern conductor (DEC). The red outline around the DDC structure denotes the perturbed area in the sensitivity test. c) Structural interpretation from a seismic profile close to our MT line (Aragón Arreola, 2006). Vertical axis is two-way travel time. The approximate depth range covered by this seismic section is indicated by the dashed line of the section in a).

230 231

The main features of the model are indicated in Figure 4b as shallow conductor

232

(SC), shallow resistor (SR), deep resistor (DR), deep central conductor (DCC). The

233

shallow conductor (SC) is associated with the high conductivities of the basin

234

sediments, with resistivities from 0.3 to 3.3 Ωm. This conductor has variable

235

thicknesses, reaching values of 5.5 km. The depths of maximum positive vertical

236

resistivity gradients are not at the base of the basin but at shallower depths of about 3

237

km, possibly associated with a decrease in the porosity of the sediments. Between sites

238

7 and 8 the conductor completely disappears, being replaced by a resistive zone, here

239

named as the shallow resistor (SR). There is no evidence in the seismic section (Figure

240

4c) that under this region the acoustic basement is closer to the surface to produce a

241

resistive zone. We believe the presence of this resistive body is artificial because the

242

inversion program tends to create shallow resistive bodies when the separation

243

between two adjacent sites is large. A similar situation, although less pronounced,

244

occurs at shallow depths between sites 8 and 9.

245 246

Lewis et al. (2001) estimated, using the receiver function approach, crustal

247

thicknesses at this latitude of 18 and 15 km at the western and eastern shores of the

248

Gulf of California, respectively. At depths corresponding to the upper mantle, the

249

electrical resistivity structure in this model is asymmetrical. In the western part of the

250

model, belonging to the Pacific Plate, the upper mantle is resistive, illustrated by the

251

deep resistor (DR) zone of Figure 4b. In contrast, the eastern part of the inverted model,

252

which belongs to the North American Plate, is more conductive.

253 254

The most interesting geophysical feature in the model of Figure 4b is the

255

presence of the deep central conductor (DCC). It has resistivities from 2 to 6 Ωm and is

256

surrounded by a more resistive host. It is located under site 5, in the northeastern limit

257

of the Wagner basin, extending from the base of the basin to greater depths. The

258

geometry of this conductor shows a dip toward the NE; at a depth of 15 km apparently

259

displays a horizontal branch toward sites 8 and 9. However, inferences about the

260

deepest structure should be taken with caution as 3D effects, revealed by the long-

261

period impedance polar diagrams of Figure 2, since they are not considered in the 2D

262

inversion approach. We interpret the low resistivity, location, and depth extension of this

263

conductive body as evidence of the presence of partial melt at the boundary between

264

the Pacific and North American Plates.

265

With the purpose of verifying the presence of the DCC anomaly, we carried out

266

two testing approaches. In the first one, we increased the resistivities of the outlined

267

area framing the DCC anomaly in Figure 4b to those of the surrounding host (12 and 24

268

Ωm), in such a way that the anomalous character of the DCC feature disappears. We

269

then calculated the forward response of this perturbed model and compared it with the

270

calculated response from the inverted model. If there is no difference between both

271

responses means that the DCC is not well constrained by the data. However, if there is

272

a significant difference means that this resistivity feature is well constrained by the

273

measured data. We used the normalized residuals as a measure to quantify the

274

response differences, defined by

275

resistivity at a given period and

276

period of the inverted model or the perturbed model, both logarithmically defined. The

277

residuals in Site 5 for eight periods, from 500 to 6,000 seconds, for the TE and TM

278

models are shown in Figure 5. Site 5 is directly above the DCC anomaly. For each

279

period the left bar corresponds to the residual with the inverted or chosen model, the

280

right bar is for the perturbed model. The elimination of the DCC results in a slight

281

decrease in the perturbed residuals. However, the residuals are significantly higher for

282

the TM mode, indicating that the DCC feature is well constrained by the data.

283 284

, where

is the observed apparent

is the calculated apparent resistivity at the same

285 286

Figure 5. Residuals from the inverted (chosen) and perturbed models at Site 5 for

287

several long-period data. (a). TM mode, (b) TE mode.

288

An additional methodology was applied to support the existence of the DCC. To

289

this end, we use the approximate formula

290

where

291

the depths of penetration

292

,

represents the average of conductivity over a depth window bounded by and

. Apparent conductivity

,

corresponds to the period

, and

to

, with

stands for the reciprocal of apparent resistivity

.

293

Unlike models that fit the data, the averages are unique and have a meaningful variance

294

(Gómez-Treviño, 1996). The wider the depth window is, the smaller the variance and

295

vice versa. As

296

(Jones, 1983), and the variance increases without bound. Different averages with their

297

corresponding variances can be computed at will, just by increasing the steps between

298

periods. The results are typical Backus-Gilbert trade-offs between variance and

299

resolution.

300

determinant of the impedance tensor for station 5 and its four neighbors, stations 3, 4,

301

6, and 7. We assumed the same error floor of 10% as we did for the 2D inversion.

the formula reduces to the familiar Niblett-Bostick transformation

We applied this approach to apparent conductivities derived from the

302 303

The results are shown in Figure 6 for the five contiguous stations. Consider first

304

station 5 and follow the averages and their variances as they develop with increasing

305

depth. It can be observed that there is definitively a decrease in resistivity below 9 km

306

since a horizontal line cannot be drawn within the region defined by the error bars. This

307

complies with the view at depth below station 5 as shown in Figure 4.

308

309 310

Figure 6. Depth averages of electrical resistivity calculated with the method proposed by

311

Gómez-Treviño (1996). The results have shown that station 5 is sensible to lower

312

resistivities, which are related to the presence of a conductive body beneath the depth

313

of 9 km.

314

315

Figure 7 shows the location of the heat flow measurements in the area (Henyey

316

and Bischoff, 1973; Prol-Ledesma et al., 2013; Neumann et al., 2017), together with the

317

mapped faults interpreted from seismic reflection profiles (Martín-Barajas et al., 2013)

318

and the MT sites. Notable features of the heat flow measurements are their high values

319

and their strong spatial variability, indicators of the presence of an intense system of

320

hydrothermal circulation, occurring mainly along faults and fractures (Prol-Ledesma et

321

al., 2013; Neumann et al., 2017). Most of the higher heat flow values occur in the

322

vicinity of the Wagner fault and smaller faults that mark the eastern boundary of the

323

basin. According to the seismic reflection information, most of the current differential

324

movement is occurring along this fault (Aragón-Arreola and Martín-Barajas, 2007).

325

Persaud et al. (2003), in an independent seismic reflection survey of the area, also have

326

found that at the southeastern side of the basin the faults are more active and the fault

327

density is higher. It is interesting to note that the three MT soundings (sites 5, 6, and 7)

328

that give the lowest resistivities at shallow levels are located close to this zone of high

329

heat flow values, strongly suggesting the presence of a correlation between low

330

resistivities and high heat flows. Figure 8 presents this correlation. In the x-axis are the

331

most superficial resistivities of six MT sites (Table 1), which are practically the same as

332

the apparent resistivities measured at the shortest period of 0.13 s (Figure 3); the

333

remaining four MT sites are not considered in this analysis because they do not have

334

any nearby heat flow measurements. In the y-axis are the average values estimated

335

from 27 heat flows measured by Neumann et al. (2017), considering in this estimation

336

only heat flow points located less than 4 km from the MT sites. The standard deviations

337

are displayed as error bars. The correlation between low resistivities and high heat

338

flows is clearly high. A controlled-source electromagnetic study (Reyes Ortega, 2016)

339

also produced lower resistivities in the same region of high heat flows.

340 341

Figure 7. Color dots indicate the locations and intensities of heat flow measurements

342

reported by Henyey and Bischoff (1973), Prol-Ledesma et al. (2013), and Neumann et

343

al. (2017). The colored bar at right is the logarithmic heat flow scale (in mW/m2). Black

344

dots denote the MT sites (the inset indicates the MT site numbers). The fault pattern is

345

from Martín-Barajas et al. (2013).

346 347 348

349 350 351

Figure 8. Mean heat flow values against the estimated shallow resistivities under six MT sites. The standard deviations are represented by bars.

352 353

The electrical conductivity in geothermal environments may be produced not only by

354

high temperatures, but also by a number of other factors. A commonly used relation for

355

the bulk conductivity

356

357

of a porous rock is given by Ward (1990),

, where a and m are factors depending on the rock texture,

is the fractional porosity,

358

is the liquid saturation,

is the fluid conductivity, and

represents an increase in

359

the bulk conductivity produced by the presence of clays. The first term is the familiar

360

Archie’s law. The second term is produced by the surface conductivity due to ions in the

361

double layer at the surface of clay particles, which also depends on several factors as

362

the clay content and the cation exchange capacity of the particular clay mineral.

363

In this work, we are interested in the relationship between the conductivities of

364

the rock and the fluid contained in their pores. The fluid conductivity essentially depends

365

on the amount of dissolved salts and temperature. To that aim, in the above expression

366

we adopt three approximations for the conductivity of the shallow sediments deposited

367

in the Wagner basin: a) these sediments are fully saturated, that is,

368

assume a fractional porosity of 0.5, and a and m parameters of 1 and 2, respectively,

369

which are reasonable assumptions (Nafe and Drake, 1957), and c) the contribution to

370

the bulk conductivity from the surface conductivity is negligible. The effect of clays in

371

unsaturated or saturated rocks with freshwater may be significant. However, when the

372

saturation fluid has high salinities, such as seawater, its effect is small (Boadu and

373

Seabrook, 2006). Under these considerations the relation between the rock resistivity

374

and the fluid resistivity is

, b) we

.

375

The resulting fluid resistivities of the six MT sites of Figure 8 are listed in Table 1.

376

We then used the Keller and Frischknecht (1966) graphs describing the resistivity

377

variation of a sodium chloride solution with concentration and temperature (Figure 9)

378

and plotted the six pore-water resistivities in two extreme situations. In the first one, we

379

assume the resistivities vary only with the temperature and not by the effect of the

380

salinity, so we plot them along the vertical line corresponding to the typical seawater

381

salinity of 35 g/kg or 35,000 ppm. In an independent way, we calculated the

382

temperature of seawater from the fluid resistivity using the relationship proposed by

383

Constable et al. (2009). At sites 3 and 6 the calculated temperatures are 18o C and 58o

384

C, respectively, agreeing well with the curves drawn in Figure 9. These temperatures

385

are listed in Table 1. In the second situation, we assume the resistivity variation is not

386

due to temperature but only to changes in the pore-water salt concentration. The points

387

are now plotted (Figure 9) along the curve of 15 oC, which is approximately the

388

temperature measured at the bottom of the water column (Neuman et al., 2017). The

389

resistivity of the pore-water under site 6 could be produced by a concentration of 200

390

g/kg, a salinity almost six higher than that of marine water (Table 1).

391 392

MT site

Rock resistivity (Ωm)

Fluid resistivity (Ωm)

Temperature (°C) at a 35 g/kg salinity

2 3 4 5 6 7

0.465 0.878 0.672 0.593 0.311 0.472

0.116 0.22 0.168 0.148 0.078 0.118

61 18 35 40 92 60

Salinity (g/kg) at a 15oC temperature 95 39 56 67 200 93

393 394

Table 1. Shallow rock resistivities interpreted with the MT data, pore-water resistivities

395

estimated with Archie’s law, temperatures at a salinity of 35 g/kg, and salinities at a

396

temperature of 15 oC under six MT sites.

397 398

We cannot know how much the temperature and the salinity contribute to the

399

anomalous pore-water resistivities; any combination of these variables enclosed within

400

the triangular area bounded by the plotted points of Figure 9 is possible. However, we

401

could attempt to constrain the salinities. With this goal, we used the brine

402

concentrations reported by Harthill (1978) of five geothermal fields located to the north

403

of the Wagner basin: four in the Imperial Valley, USA (Salton Sea, Burec, Wilson, and

404

Heber) and one in the Mexicali Valley, Mexico (Cerro Prieto). All these fields are in the

405

region of continental aperture at the boundary between the North America and Pacific

406

plates. In Figure 9, the sum of these concentrations plus that of seawater is plotted. The

407

lower (48 g/kg) and upper (164 g/kg) bounds of these concentrations could serve to limit

408

the

Ω 409 410 411 412 413 414 415 416

Figure 9. Resistivity of a sodium chloride solution as a function of temperature and salt concentration (after Keller and Frischknecht, 1966). The fluid resistivity estimates from the six Wagner basin MT sites are plotted assuming a constant salinity and a constant temperature. The salinity ranges from five geothermal fields in the Mexicali and Imperial Valleys (Harthill, 1978) are indicated.

417

possible combination of temperature and salinity that could produce such resistivities.

418

Another source of information on the salinities is the chemical analysis of pore-water

419

samples taken from the first three meters of sediments (R. Batista, pers. comm.). In a

420

location 3 km northward of our MT site 6, the mean salinity of 11 samples is 37.1 g/kg

421

with a standard deviation of 3.8 g/kg. At this MT site, a low resistivity and high heat flow

422

were measured (Figure 8), suggesting that the anomalously low resistivities might be

423

due mainly to high temperatures, and not to high salinities.

424 425

Discussion

426

We interpret the deep central conductor as a zone of partial melt and the ultimate

427

source of heat for the surface hydrothermal upwelling. It is the first time such structure

428

has been sensed in the Gulf of California using an electromagnetic method. The

429

conductor shows an apparent dip toward the East, probably related to the Wagner fault

430

(González-Escobar et al., 2009), which marks the eastern boundary of the depocenter

431

and is tectonically more active than the western limit of the basin (Aragón-Arreola and

432

Martín-Barajas, 2007). Information supporting the presence of magmatic intrusions at

433

the base of the sedimentary deposits, which is contributing to the accretion of oceanic

434

crust, comes from two sources. From the Bouguer anomaly (shown in Figure 4a)

435

measured along a line almost coincident with our MT profile by Petróleos Mexicanos

436

(Pemex), Pérez (1982) has interpreted a high local gravity as produced by mafic

437

intrusions at the base of the basin. Furthermore, in a seismic reflection line about 15 km

438

to the south of our profile, over the Consag basin González et al. (2014) detected a

439

reflection interpreted as a volcanic intrusion at a depth of 6 or 7 km.

440 441

Considering the resistivity of this conductor (4 Ωm) as the effective value of a

442

two-component medium we can have an estimate of the amount of melt in this mixture.

443

These two components are magma, with a 0.25 Ωm resistivity (Constable, 2007), and

444

the 30 Ωm normal resistivity surrounding the conductor in the model of Figures 4b.

445

Using the bounds proposed by Hashin-Shtrikman (1962) for a two-component system,

446

the bounds for the melt fraction vary from 8% to 70%, which is a range too wide to be

447

useful. If we now consider Archie’s law with a cementation factor of 2, we obtain an

448

estimate of 25 % melt in this conductor.

449 450

There have been several MT studies over mid-ocean spreading ridges where the

451

conductor associated with the zone of partial melt is vertical (Heinson et al., 2000; Key

452

et al., 2013), asymmetric (Baba et al., 2006a) or even absent (Baba et al., 2006b). Its

453

presence depends on such factors as pressure, temperature, water content, and

454

connectivity between the melt pockets.

455 456

A number of studies describing the shear velocity structure of the Gulf area

457

based on Rayleigh wave dispersion have been carried out (e.g., Zhang et al., 2007;

458

Wang et al., 2009; Di Luccio et al., 2014). They show complex variations of the velocity

459

in both horizontal and vertical directions. In none of them, there is an agreement

460

between our deep central conductor and their low-velocity zones in the depth range

461

from 10 to 40 km.

462

horizontal resolution of 10 by 10 km of some of these seismic studies (Di Luccio et al.,

463

2014), which might not be good enough to resolve the 10 to 15 km width of the

464

conductor.

One possible reason for this discrepancy is the approximate

465 466

Conclusions

467

The inverted 2-D model using the data from both polarizations gave responses

468

with adequate fits with the observed data. The main features of this model are the low

469

resistivities associated with the sedimentary basin and a deep conductor. We obtained

470

a positive correlation between the resistivities estimated in the first 100 m and the

471

measured heat flow values. Most of the higher heat flows occur in the vicinity of the

472

Wagner fault and smaller associated faults, where present-day tectonic activity is

473

inferred from seismic reflection data. The two MT sites with the lowest shallow

474

resistivities are located in this zone. Although low resistivities might also be produced by

475

high salinities of the hydrothermal plumes, direct measurements of the salt content of a

476

few shallow pore-water samples might suggest that the anomalous resistivities are

477

mainly influenced by the high temperatures. The DCC is interpreted as a zone of partial

478

melt feeding the formation of new crust. It has a resistivity close to 4 Ωm, extends from

479

the base of the shallow conductor to 15 km depths, and has an apparent dip toward the

480

East. This dipping is probably related to the higher tectonic activity of the eastern side of

481

the basin. Assuming that its effective resistivity follows Archie’s law, we estimate a 25 %

482

melt in this conductor.

483 484

Acknowledgments

485

This project was funded by CeMIEGeo and CICESE. We acknowledge the captain and

486

crew of Alpha Helix, the CICESE research vessel, for their help during the data

487

acquisition stage. We thank David Myer, Ramón Batista, Guillermo Díaz, Mario

488

González, Arturo Martín, and Martín Pacheco for their technical advice in different

489

aspects of this work. T.A.E. and V.R.O. were supported by scholarships from CONACyT

490

and by UC-Mexus for a temporal stay at Scripps.

491 492

References

493

Aragón Arreola, M., 2006, Structural evolution of basins in the northern and central Gulf

494

of California. Implications for rift kinematics and strain accommodation, Ph.D. Thesis,

495

Centro de Investigación Científica y Educación Superior de Ensenada, Baja California,

496

Ciencias de la Tierra, 118 pp.

497 498

Aragón-Arreola, M., and Martín-Barajas, A. 2007, Westward migration of extension in

499

the northern Gulf of California, Mexico, Geology, 35(6), 571–574.

500 501

Atwater, T., 1970, Implications of Plate Tectonics for the Cenozoic tectonic evolution of

502

western North America, Geol. Soc. Am. Bull., 81, 3513-3536.

503 504

Avilés Esquivel, T., 2016, A Magnetotelluric study at the North of the Gulf of California

505

(in Spanish), M.Sc. Thesis, Earth Sciences Division, CICESE, Mexico, 67 pp.

506 507

Baba, K., Chave, A. D., Evans, R. L., Hirth, G., and Mackie, R. L. , 2006a, Mantle

508

dynamics beneath the East Pacific Rise at 17o S: Insights from the Mantle

509

Electromagnetic and Tomography (MELT) experiment, J. Geophys. Res., Solid

510

Earth, 111, B02101.

511 512

Baba, K., Tarits, P., Chave, A. D., Evans, R. L., Hirth, G., and Mackie, R. L. , 2006b,

513

Electrical structure beneath the northern MELT line on the East Pacific Rise at 15o 45’

514

S, Geophys. Res. Lett., 33, L22301.

515 516

Batista Cruz, R.Y., Peiffer, L., Weber, B., Herguera, J.C., Liotta, M., González

517

Fernández, A., Barragán Reyes, R.M., and Kretzschmar, T.G., 2019, Geochemical

518

characteristics of pore waters from sediment cores of the Wagner Basin, Gulf of

519

California, Applied Geochemistry, 13 pp.

520 521

Boadu, F.K., and Seabrook, B.C., 2006, Effects of clay content and salinity on the

522

spectral electrical response of soils, J. of Environmental and Engineering Geophysics,

523

11(3), 161-170

524 525

Boyd, T.L., Sifford, A., and Lund, J.W., 2015, The United States of America country

526

update 2015, Proc. World Geothermal Cong. 2015, Melbourne, Australia, 12 pp.

527 528

Canet, C., Prol-Ledesma, R.M., Dando, P.R., Vázquez-Figueroa, V., Shumilin, E.,

529

Birosta, E., Sánchez, A., Robinson, C.J., Camprubí, A., and Tauler, E., 2010, Discovery

530

of massive seafloor gas seepage along the Wagner Fault, northern Gulf of California,

531

Sedimentary Geology, 228, 292-303.

532 533

Constable, S. C., Parker, R. L., and Constable, C. G. ,1987, Occam’s inversion: A

534

practical algorithm for generating smooth models from electromagnetic sounding

535

data, Geophysics, 52(3), 289-300.

536

537

Constable, S.C., Heinson, G.S., Anderson, G., and White, A., 1997, Seafloor

538

electromagnetic measurements above Axial Seamount, Juan de Fuca Ridge, J.

539

Geomag. Geoelectr., 49, 1327-1342.

540 541

Constable, S. C., Orange, A. S., Hoversten, G. M., and Morrison, H. F., 1998, Marine

542

magnetotellurics

543

system, Geophysics, 63(3), 816-825.

for

petroleum

exploration

Part

I:

A

sea-floor

equipment

544 545

Constable, S., 2007, Geomagnetic Induction Studies, in M. Kono (ed.), Geomagnetism,

546

Treatise on Geophysics, vol. 5, 237-276, Elsevier.

547 548

Constable, S., Key, K., and Lewis, L., 2009, Mapping offshore sedimentary structure

549

using electromagnetic methods and terrain effects in marine magnetotelluric

550

data, Geophys. J. Int., 176(2), 431-442.

551 552

Di Luccio, F., Persaud, P., and Clayton, R.W., 2014, Seismic structure beneath the Gulf

553

of California: A contribution from group velocity measurements, Geophys. J. Int., 199,

554

1861-1877.

555 556

Dorsey, R.J., 2010, Sedimentation and crustal recycling along an active oblique-rift

557

margin: Salton Trough and northern Gulf of California, Geology, 38(5), 443-446.

558 559

Egbert, G.D., 1997, Robust multiple-station magnetotelluric data processing, Geophys.

560

J. Int., 130, 475-496.

561 562

Evans, R.L., Constable, S.C., Sinha, M.C., Cox, C.S., and Unsworth, M.J., 1991, Upper

563

crustal resistivity structure of the East Pacific Rise near 13o N, Geophys. Res. Lett., 18,

564

1917-1920.

565 566

Gómez-Treviño, E., 1996, Approximate depth averages of electrical conductivity from

567

surface magnetotelluric data, Geophys. J. Int., 127(3): 762–772.

568 569

González-Escobar, M., Aguilar-Campos, C., Suárez-Vidal, F., and Martín-Barajas, A.,

570

2009, Geometry of the Wagner basin, upper Gulf of California based on seismic

571

reflections. International Geology Review, 51(2): 133–144.

572 573

González-Escobar, M., Suárez-Vidal, F., Hernández-Pérez, J.A., and Martín-Barajas,

574

A., 2010, Seismic reflection-based evidence of a transfer zone between the Wagner and

575

Consag basins: implications for defining the structural geometry of the northern Gulf of

576

California, Geo-Mar. Lett., 30, 575-584.

577 578

González-Escobar, M., Suárez-Vidal, F., Sojo-Amezquita, A., Gallardo-Mata, C. G., and

579

Martín-Barajas, A., 2014, Consag basin: northern Gulf of California, evidence of

580

generation of new crust, based on seismic reflection data. International Geology

581

Review, 56(11), 1315–1331.

582 583

Grijalva, N., 1986, Investigación de la energía geotérmica en la depresión de Wagner

584

en el Golfo de California, latitud 31º 00’ al 31º 15’ y longitud 113º 50’ al 114º 10’ ,

585

Unpublished report to the Comisión Federal de Electricidad, Cerro Prieto, B.C.

586 587

Gutiérrez-Negrín, L.C.A., Maya-González, R., and Quijano-León, J.L., 2015, Present

588

situation and perspectives of geothermal in Mexico, Proc. World Geothermal Cong.

589

2015, Melbourne, Australia, 10 pp.

590 591

Harthill, N., 1978, A quadripole resistivity survey of the Imperial Valley, California,

592

Geophysics, 43, 1485-1500.

593 594

Heinson, G., Constable, S., and White, A., 2000, Episodic melt transport at mid-ocean

595

ridges inferred from magnetotelluric sounding, Geophys. Res. Lett., 27, 2317-2320.

596 597

Hashin, Z., and Shtrikman, S., 1962, A variational approach to the theory of the effective

598

magnetic permeability of multiphase materials, J. of Applied Phys., 33(10), 3125-3131.

599 600

Henyey, T.L., and Bischoff, J.L., 1973, Tectonic elements in the northern part of the Gulf

601

of California, Geol. Soc. Am. Bull., 84, 315-330.

602 603

Jones, A.G., 1983, On the equivalence of the “Niblett” and “Bostick” transformations in

604

the magnetotelluric method, J. Geophys., 53, 72-73.

605 606

Key, K., and Weiss, C., 2006,. Adaptive finite-element modeling using unstructured

607

grids: The 2D magnetotelluric example. Geophysics, 71(6), G291-G299.

608 609

Key, K., and Ovall, J., 2011, A parallel goal-oriented adaptive finite element method for

610

2.5-D electromagnetic modeling, Geophys. J. Int., 186(1), 137-154.

611 612

Key, K., 2012, Marine electromagnetic studies of seafloor resources and tectonics,

613

Surv. Geophys., 33, 135-167.

614 615

Key, K., Constable, S., Liu, L., and Pommier, A. , 2013, Electrical image of passive

616

mantle upwelling beneath the northern East Pacific Rise, Nature, 495, 499-502.

617 618

Key, K. , 2016, MARE2DEM: a 2-D inversion code for controlled-source electromagnetic

619

and magnetotelluric data, Geophys. J. Int., 207(1), 571-588.

620 621

Klitgord, K., Mudie, J., Bischoff, J., and Henyey, T., 1974, Magnetic anomalies in the

622

northern and central Gulf of California, Geol. Soc. of America Bull., 85(4), 815–820.

623 624

Larson, R. L., Menard, H. W., and Smith, S. M., 1968, Gulf of California: A result of

625

ocean-floor spreading and transform faulting, Science, 161(3843), 781-784.

626 627

Lewis, J.L., Day, S.M., Magistrale, H., Castro, R.R., Rebollar, C., Astiz, L., Eakins, J.,

628

Vernon, F., and Brune, J.N., 2001, Crustal thickness of the Peninsular Ranges and Gulf

629

Extensional Province in the Californias, J. Geophys. Res., 106, B7, 13,599-13,611.

630 631

Lonsdale, P., 1989, Geology and tectonic history of the Gulf of California, in E.L.

632

Winterer, D.M. Hussong, and R.W. Decker (eds.), The Geology of North America, vol.

633

N, The eastern Pacific Ocean and Hawaii, The Geological Society of America, 499–521.

634 635

Martín-Barajas, A., González-Escobar, M., Fletcher, J.M., Pacheco, M., Oskin, M., and

636

Dorsey, R., 2013, Thick deltaic sedimentation and detachment faulting delay the onset

637

of continental rupture in the northern Gulf of California: Analysis of seismic reflection

638

profiles, Tectonics, 32, 1294-1311.

639 640

Nafe, J.E., and Drake, C.L., 1957, Variation with depth in shallow and deep water

641

marine sediments of porosity, density and the velocities of compressional and shear

642

waves, Geophysics, 22(3), 523-552.

643 644

Nagy, E.A., and Stock, J.M., 2000, Structural controls on the continent-ocean transition

645

in the northern Gulf of California, J. Geophys. Res., 105(B7), 16251-16269.

646 647

Neumann, F., Negrete-Aranda, R., Harris, R.N., Contreras, J., Sclater, J.G., and

648

González-Fernández, A., 2017, Systematic heat flow measurements across the Wagner

649

Basin, northern Gulf of California, Earth and Planet. Sci. Ltts., 479, 340-353.

650 651

Pérez, C. G. 1982. Algunos resultados de la investigación geológico-geofísica en la

652

porción noroccidental del Golfo de California, Asoc. Mexicana de Geofísicos de

653

Exploración, 34, 71–77.

654 655

Persaud, P., Stock, J.M., Stecler, M.S., Martín-Barajas, A., Diebold, J.B., González-

656

Fernández, A., and Mountain, G.S., 2003, Active deformation and shallow structure of

657

the Wagner, Consag, and Delfín Basins, northern Gulf of California, Mexico, J.

658

Geophys. Res., 108, B7, 2355.

659 660

Prol-Ledesma, R. M., Torres-Vera, M.-A., Rodolfo-Metalpa, R., Angeles, C., Lechuga

661

Deveze, C. H., Villanueva-Estrada, R. E., Shumilin, E., and Robinson, C., 2013, High

662

heat flow and ocean acidification at a nascent rift in the northern Gulf of California,

663

Nature Communications, 4, 1388.

664 665

Reddy, I.K., Rankin, D., and Phillips, R.J., 1977, Three-dimensional modeling in

666

magnetotelluric and magnetic variational sounding, Geophys. J.R. astr. Soc., 51, 313-

667

325.

668 669

Reyes Ortega, V., 2016, Interpretation of marine electromagnetic data of the Wagner

670

Basin, M.Sc. Thesis, CICESE, Ensenada, Mexico, 82 pp.

671 672

Sánchez-Zamora, O., Doguin, P., Couch, W., and Ness, G., 1991, Magnetic anomalies

673

of the northern Gulf of California: Structural and thermal interpretations, in Dauphin,

674

J.P., and Simoneit, B.R.T., (eds.), The Gulf and Peninsular Province of the Californias,

675

Am. Ass. Petrol. Geol., Memoir 47, 377-401.

676 677

Wang, Y., Forsyth, D. W., and Savage, B., 2009, Convective upwelling in the mantle

678

beneath the Gulf of California, Nature, 462, 499-501.

679 680

Ward, S.H., 1990, Resistivity and Induced Polarization Methods, in Ward, S.H. (ed.),

681

Geotechnical and Environmental Geophysics, Vol. I: Review and tutorial, 147-189, Soc.

682

of Explor. Geophys., Tulsa, Oklahoma.

683 684

Zhang, X., Paulssen, H., Lebedev, S., and Meier, T., 2007, Surface wave tomography of

685

the Gulf of California, Geophys. Res. Lett., 34, L15305.

686

Highlights • • •

The magnetotelluric marine method was used to explore the North Gulf of California with geothermal purposes. The 2D model of the magnetotelluric marine data reveals a deep central conductor in the center of the Wagner basin. We obtained a positive correlation between the low resistivities estimated in the first 100 m using 2D inversion of the magnetotelluric marine data with high heat flow measurements reported previously.

January 10th, 2020 Declaration of interests ☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: