*Manuscript Click here to download Manuscript: manuscript.docx Click here to view linked References

1 Southeastern migration of the Laramide porphyry copper mineralization in 2 northeastern based on available and new Re-Os and U-Pb geochronology: 3 implications in exploration along the Cananea lineament. 4 5 6 Rafael Del Rio-Salas (1); Lucas Ochoa-Landín (2); Martín Valencia-Moreno (1); Thierry 7 Calmus (1); Diana Meza-Figueroa (2); Sergio Salgado-Souto (3,4); Jason Kirk (3); Joaquin 8 Ruiz (3); Héctor Mendívil-Quijada (5) 9 10 (1) Estación Regional del Noroeste, Instituto de Geología, Universidad Nacional Autónoma 11 de Mexico. L.D. Colosio y Madrid S/N, Col. Los Arcos, Hermosillo, Sonora, 83240 México. 12 13 (2) Departamento de Geología, División de Ciencias Exactas y Naturales, Universidad de 14 Sonora, Rosales y Encinas, Hermosillo, Sonora, 83000 México. 15 16 (3) Geosciences Department, University of , 1040 E 4th St, Tucson Arizona, 85721 17 USA. 18 19 (4) U.A. Ciencias de la Tierra, Universidad Autónoma de Guerrero, Ex-Hacienda San Juan 20 Bautista s/n, Taxco el Viejo, Guerrero 40323, México 21 22 (5) GEO Digital Imaging de México, S.A. de C.V, Hermosillo, Sonora, México. 23 24 25 26 Corresponding Author: [email protected] 27 28 29 30 31 32 33 34 35 36 37 To be submmited to Ore Geology Reviews 38 39 40 41 42 43 44 45 46 47 ABSTRACT

48 The Cananea mining district encompasses the most important copper deposits in Mexico, which,

49 together with the porphyry copper deposits from southern Arizona and western New Mexico,

50 conform one of the largest copper provinces on Earth. The main mineralized localities within the

51 district are the Buenavista del Cobre, formerly called the Cananea mine, Mariquita, Milpillas,

52 Lucy, María, El Alacrán, and El Pilar deposits. The Buenavista del Cobre, is far the most

53 relevant site with more than 30 Mt Cu, which places it among a selected group of giant porphyry

54 copper deposits at world scale. The present paper addresses new Re-Os molybdenite and U-Pb

55 zircon ages, which help to better constrain the metallogenetic history of the district. According to

56 the Re-Os data, the earlier mineralizing pulse occurred at the El Pilar (73.9 Ma), followed by the

57 Milpillas (63.0 to 63.1 Ma), Lucy (61.6 to 61.8 Ma), María (60.4 Ma), El Alacrán (60.8 to 60.9

58 Ma), Buenavista del Cobre (59.2 to 59.3 Ma), and Mariquita (59.3 ± 0.3 Ma) deposits. Excluding

59 El Pilar, the main mineralizing pulses occurred within a relatively short time window of ~4 Ma.

60 Regarding the magmatic history, U-Pb zircon ages for the emplacement of the copper-related

61 plutons, including the El Pilar deposit, which lacks an exposed porphyritic stock, the entire

62 magmatic activity occurred from 77.7 to 57.8 Ma. The Re-Os and U-Pb geochronological data

63 indicate a NW to SE progression of the mineralizing events along the trace of the so-called

64 Cananea lineament, suggesting a regional structural control coupled with southeastward

65 migration of the magmatism. Thus, our data show a clear periodicity in the magmatic-

66 hydrothermal events responsible for the emplacement of the porphyry copper ores in the

67 Cananea district. Tertiary extension dismembered and buried part of the porphyry copper

68 systems, a fact that opens an interesting opportunity to explore dissected, tilted, and hidden parts

69 of the ore systems. Moreover, Tertiary extension promoted formation of important supergene 70 enriched chalcocite blankets, and lateral migration of copper solutions, yielding conditions to

71 develop exotic copper mineralization.

72

73

74 KEYWORDS: Porphyry copper deposits; Re-Os molybdenite age; U-Pb zircon age; Cananea

75 district

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92 93

94 1. INTRODUCTION

95 Late Cretaceous-early Paleogene Laramide porphyry copper deposits are

96 widespread in northwestern Mexico, among which the Buenavista del Cobre (before the

97 Cananea mine) and La Caridad deposits (Fig. 1) stand as the most remarkable examples,

98 comprising the largest national copper resources, being also important at world-scale,

99 accounting for more than 40 Mt of Cu (Valencia-Moreno et al., 2006, 2007).

100 The Cananea and Nacozari mining districts extend as the southern part of the

101 porphyry copper province of North American Cordillera (Titley, 1982), also referred to as

102 the “great cluster” of the porphyry copper deposits (Keith and Swan, 1996). These mining

103 districts also contain smaller porphyry copper deposits (PCD) and other metal

104 commodities and mineralization styles, including skarn, manto, and hydrothermal breccias.

105 The Cananea region has been target of several geological studies (e.g., Aponte-

106 Barrera, 2009; Carreón-Pallares, 2002; Mulchay and Velasco, 1954; Valentine, 1936;

107 Velasco, 1966), magmatic, hydrothermal, and mineralization studies (Bushnell, 1988;

108 Meinert, 1982, Varela, 1972; Wodzicki, 2001), geochemical studies focused on

109 characterizing the magmatic evolution (Wodzicki, 1995, 2001), supergene enrichment

110 processes studies (Noguez-Alcántara et al., 2007; Virtue, 1996), and fluid inclusions

111 microthermometric studies (Arellano-Morales, 2004; González-Partida et al., 2013).

112 Previous geochronological data help to constrain the age of the different magmatic

113 pulses and mineralization episodes recorded in the Cananea and Nacozari mining districts

114 (Anderson and Silver, 1977; Barra et al., 2005; Damon et al. 1983; Noguez-Alcántara et al.,

115 2007; Valencia et al., 2006). However, considering the economic importance, not only at 116 district level, but also at world-scale, the available data are not enough to have complete

117 picture of the magmatic and hydrothermal histories of the porphyry copper systems in

118 northeastern Sonora. This paper contributes with new U-Pb zircon and Re-Os molybdenite

119 geochronological data from the Cananea mining district, to better constrain the timing and

120 evolution of the different magmatic and mineralizing pulses that culminated with the

121 emplacement of world-class copper deposits, which makes this region highly attractive for

122 further exploration campaigns. In addition, our data may provide new clues to determine

123 the spatial distribution and possible links between the Cananea and Nacozari districts.

124

125 2. REGIONAL GEOLOGICAL FRAMEWORK

126 The geology of northwestern Mexico records an important part of the tectonic

127 history of the western North American Cordillera. Available geological, geochemical, and

128 isotopic data allow reconstruction of a series of continental magmatic arcs built on

129 Proterozoic crystalline basement. The older Cordilleran arc is portrayed by scattered

130 plutonic and volcanic rocks of Late Triassic to Late Jurassic age (Anderson and Silver, 1979;

131 Busby-Spera, 1988; Grajales-Nishimura et al., 1992), which were emplaced well inside the

132 continent (Fig. 1). In Early Cretaceous, a new magmatic activity took place in a region closer

133 to the paleotrench, developing a well-defined arc that extended along Baja California and

134 westernmost Sonora and Sinaloa, which is best illustrated by the Peninsular Ranges

135 batholith. The igneous activity was nearly stationary during most of the Cretaceous time,

136 from ~139 to 80 Ma (Henry et al., 2003; Silver and Chappell, 1983); however, regional

137 chrontours show a clear eastward migration of the igneous activity (Ortega-Rivera, 2003;

138 Valencia-Moreno et al., 2006). 139 After 80 Ma, the locus of magmatism migrated faster to the east, reaching eastern

140 Sonora ~59 Ma ago, although reconstructed chrontours indicate a much slower migration

141 rate to the south, across southern Sonora and Sinaloa (Valencia-Moreno et al., 2006). The

142 accelerated stage of arc migration is currently associated with the Laramide Orogeny (80-

143 40 Ma), which was triggered by a rapid increase in the rate of the Farallon-North America

144 plate convergence, in response to major global tectonic adjustments (e.g, Coney and

145 Reynolds, 1977; Damon et al., 1983). The Laramide Orogeny caused crustal shortening and

146 widespread magmatism in most of the Cordillera. In contrast to the Sierra Madre Oriental

147 fold and thrust belt of eastern Mexico, in northwestern Mexico, and particularly in Sonora,

148 the Laramide shortening is yet poorly constrained, and evidenced by block-uplifting

149 tectonics, although thin-skin deformation is documented in Jurassic to Late-Cretaceous

150 volcanic and sedimentary sequences (Calmus et al., 2011; Davis, 1979). The associated

151 magmatism left a broad belt of volcano-plutonic rocks, with the volcanic member best

152 preserved in the eastern part of it, whereas western part essentially exhibits large

153 composite batholiths. This fact may have been the result of post-Laramide exhumation

154 processes.

155 The Laramide igneous activity, coupled with a particular plate tectonic geometry

156 that allowed profuse asthenospheric mantle melting, yielded conditions, not only for the

157 generation of the relatively long-lived calc-alkaline magma chambers, required to form

158 these volcanoplutonic complexes, but also for the emplacement of subvolcanic magmatic-

159 hydrothermal systems. These systems controlled circulation of hydrothermal fluids that

160 concentrated tremendous amounts of metals, particularly Cu and Mo, giving place to

161 formation of an important number of porphyry copper type deposits. Laramide porphyry 162 copper deposits are known along the entire Cordillera, however, the region of

163 southwestern North America, particularly, southern Arizona, western New Mexico, and

164 northeastern Sonora (Fig. 2) concentrated the largest amount of copper, making this region

165 one of the largest Cu provinces on Earth (Titley, 1993).

166

167 2.1 CANANEA MINING DISTRICT

168 The Cananea mining district is located in a region underlain by the Precambrian

169 North American craton, within the Basin and Range extensional province (Fig. 1). The

170 Precambrian basement belongs to the Mazatzal Paleoproterozoic province (1.6-1.7 Ga),

171 extending through southern Arizona, New Mexico, and northern Sonora (Amato et al.,

172 2008). This basement is characterized in Sonora by the Pinal schist, which is intruded by

173 the 1.4 Ga Mesoproterozoic anorogenic Cananea granite (Anderson and Silver, 1977),

174 which is constituted by two magmatic facies: (1) a granophyric granitoid with phenocrysts

175 of quartz in a matrix of K-feldspar, quartz, and oligoclase; (2) less abundant coarse-grained

176 to pegmatitic granitoid composed of K-feldspar, oligoclase, quartz, and minor hornblende,

177 magnetite, and apatite (Valentine, 1936).

178 The Proterozoic assemblage is unonformably overlain by Paleozoic rocks of the

179 North American platform that include the Bolsa (Cambrian), Abrigo (Cambrian), Martín

180 (Devonian), and Escabrosa (Mississippian) Formations, and the Naco Group (Permian),

181 which were described in Meinert (1982), Mulchay and Velasco (1954), and Velasco (1966).

182 The Proterozoic and Paleozoic rocks are unconformably overlain by Mesozoic

183 volcanic rocks (Busby-Spera, 1988; Rodríguez-Castañeda and Anderson, 2011; Valentine,

184 1936), and intruded by granites of Jurassic age (Anderson and Silver, 1977). The oldest 185 volcanic rocks correspond to 1,800 m thick pile of rhyolite and andesite interbedded with

186 sandstone and quartzite of the Triassic-Jurassic Elenita Formation (Valentine, 1936;

187 Wodzicki, 1995). The 1,700 m thick Jurassic Henrietta Formation overlies the Elenita

188 Formation (Valentine, 1936), and is composed by medium to high-K, calc-alkaline, dacitic

189 to rhyolitic tuffs and flows (Wodzicki, 1995). A hornblende Ar-Ar age from the Henrietta

190 Formation yielded a minimum age of 94 Ma (Wodzicki, 1995). The intrusive counterpart of

191 Jurassic rocks within the Cananea district is the 175 Ma old El Torre syenite, which

192 intrudes both the Elenita and Henrietta Formations (Wodzicki, 1995).

193 Late Cretaceous-early Paleogene volcanic and plutonic rocks of the Laramide arc are

194 widespread in the district (Fig. 3). According to Meinert (1982), the base of the sequence is

195 the Mariquita diabase, which consists of a high-K basaltic-andesite flows, and intrusions

196 cross-cutting the rocks of the Henrietta Formation (Wodzicki, 1995). The Mesa Formation

197 overlies the Mariquita diabase. This formation is 1,500 m thick, and mostly consists of

198 intermediate volcanic rocks, which are interbedded with clastic and volcaniclastic deposits

199 (Valentine, 1936; Wodzicki, 2001). A flow within the Mesa Formation yielded a biotite Ar-

200 Ar age of 69 ± 0.2 Ma (Wodzicki, 1995). This age is confirmed by three biotite Ar-Ar ages

201 bracketed between 72.6 ± 1.2 and 65.8 ± 0.4 Ma (Cox et al., 2006).

202 The Precambrian to Cretaceous rocks are intruded by Laramide plutons and

203 porphyritic intrusions. The oldest Laramide intrusions are the Tinaja diorite and the

204 Cuitaca granodiorite (Valentine, 1936). Previous studies suggest that the Tinaja and Cuitaca

205 intrusions belong to the same batholith (Bushnell, 1988; Meinert, 1982; Valentine, 1936).

206 This is supported by isotopic data, which suggests a genetically related batholith (Wodzicki,

207 1995). Anderson and Silver (1977) obtained a U-Pb zircon age of 64 ± 3 Ma for a sample of 208 the Cuitaca batholith located near the Cananea town. A similar U-Pb zircon age of 63.8 ± 1.1

209 Ma has been reported for a sample of the Cuitaca granodiorite from Lucy pit (Fig. 3),

210 located ~20 km W-NW of Cananea (Del Rio-Salas et al., 2013).

211 Relatively small porphyritic plugs, mostly of quartz-feldspatic composition, derived

212 from the cooling batholith, are elsewhere reported within the Cananea district. These plugs

213 are regionally important since they are considered as the main heat source that centered

214 the alteration and most of the porphyry copper mineralization (Barton et al., 1995). The

215 available U-Pb zircon data for the porphyry stocks in the Cananea district indicate an age of

216 63.9 ± 1.3 Ma for a sample located in the Milpillas area (Fig. 3), ~20 km NW of Cananea

217 (Valencia et al., 2006), whereas samples from the Mariquita area, located ~14 km W-NW of

218 Cananea (Fig. 3) yielded younger U-Pb zircon ages of 62.7 ± 1.3 and 60.4 ± 1.1 (Del Rio-

219 Salas et al., 2013).

220 The Tinaja-Cuitaca batholith complex was intruded by late sub-vertical N60ºW

221 oriented mafic dikes, reported by Valentine (1936) as the Campana dikes. One sample of

222 these dikes located 2.8 km southwest of the Maria mine, yielded a hornblende Ar-Ar age of

223 58.4 ± 0.6 Ma (Carreón-Pallares, 2002).

224 The Cananea district, as well as most of the North American Cordillera, was

225 deformed by extension associated with the Tertiary Basin and Range extensional province

226 (e.g., Sonder and Jones, 1999). NNW-SSE oriented graben-horst structures alternate from

227 west to east within the district as follows: 1) the Cocóspera-San Antonio graben, and the

228 San Antonio-Chivato-Cuitaca horst; 2) the Milpillas-Cuitaca graben, and the Mariquita-El

229 Cobre-Elenita horst; and 3) the Río San Pedro-Río Sonora graben, and the Los Ajos horst.

230 During the Tertiary, the structural evolution of the Cananea district is controlled by the 231 post-Laramide Basin and Range extension. The Cananea district is located slightly east of

232 the Metamorphic Core Complex belt (Fig. 1), showing no evidence of deep extensional

233 structures, such as cataclastic fault zones or low-angle mylonitic detachment faults, which

234 suggests a low to moderate extension in the Cananea district.

235 Regarding mineralization, Cananea is the most important copper mining district in

236 Mexico, being among the world-class copper producers. Consequently, this mining district

237 has been target of several studies, particularly focused on characterizing the geology and

238 economic potential. The most important deposit within the district is by far the Buenavista

239 del Cobre mine, however there are several smaller porphyry copper, and other related

240 mineralization deposits, including breccia pipes, skarn, manto, pegmatite, polymetallic

241 veins, and replacement. In most cases, the mineralization of the Cananea mining district is

242 clearly related to magmatic pulses occurred during the Laramide arc activity. The district

243 includes several porphyry copper occurrences and related deposits (Fig. 2, Table 1),

244 currently under production (Buenavista del Cobre, Milpillas, Mariquita) or mined in the

245 past (María) or still develops sporadic mining activity (Lucy), or mining prospects (Alisos,

246 Toro). Other deposits, such as the El Alacrán and El Pilar, are considered as interesting

247 prospects (Broch, 2012; Dean, 1975, Arellano-Morales, 2004).

248

249 2.1.1 Buenavista del Cobre deposit 250 251 The Buenavista del Cobre mine, formerly Cananea, is the biggest porphyry copper

252 deposit known in Mexico (Fig. 3). The resources account for 7,140 Mt at 0.42% Cu, 0.008%

253 Mo (Singer et al., 2005). The mineralization is hosted mainly in the volcanic rocks of the

254 Henrrieta and Mesa Formations, intruded by porphyritic plugs, some of which were 255 mineralizing. The composition of these intrusions includes quartz monzonite,

256 monzodiorite, and granodiorite. Intense fracturing that developed stockwork zones

257 surround the porphyritic intrusions. The porphyry emplacement was accompanied by

258 intense hydrothermalism, characterized by potassic and quartz-sericite alteration (Ochoa-

259 Landín and Echávarri, 1978).

260 The secondary copper mineralization is nearly horizontal, extending by an area of

261 15 km2, reaching a maximum thickness of 500 m (Ayala-Fontes, 2009). Numerous

262 elliptical-shaped breccias bodies are found within the area, some of which reach depths of

263 hundreds meters, and were apparently developed during the stage of quartz-sericite

264 alteration (Bushnell, 1988). Also, associated with the porphyritic intrusions, there are some

265 skarn mineralization (Zn-Pb-Cu) and manto-type occurrences, which are hosted in the

266 intruded Paleozoic rocks, developing stratiform bodies of high-grade sulfide and iron oxide,

267 best represented by the Puertecitos skarn, located northwest of Cananea (Einaudi, 1982;

268 Meinert, 1982).

269

270 2.1.2 Milpillas deposit

271 The Milpillas porphyry copper deposit is located 20 km northwest of Cananea and 7

272 km north of Mariquita mine (Fig. 3). Milpillas mine is currently in production with 35 Mt at

273 2.3% of Cu (Noguez-Alcántara et al., 2007). Ore bodies are mostly hosted in volcanic rocks

274 of the Henrietta and Mesa Formations, and related to intrusion of porphyritic plugs, one of

275 them dated at 63.9 ± 1.3 (Valencia et al., 2006). The mineralization consists of various

276 supergene-enriched chalcocite and minor covellite blankets, overlain by Cu sulfates, Cu

277 carbonates, and Fe oxides (Noguez-Alcántara et al., 2007). Unusually for the Cananea 278 district, Milpillas developed underground mining, because the mineralization remains

279 beneath thick semi-consolidated gravel deposits of the Sonora Group (Grijalva-Noriega and

280 Roldán-Quintana, 1998).

281

282 2.1.3 Mariquita deposit

283 The Mariquita is a porphyry copper deposit located 15 km northwest of Cananea

284 (Fig. 3). The deposit is currently under production, and includes resources estimated of

285 about 58 Mt of Cu at 0.4-0.6% (Aponte-Barrera, 2009). The mineralization is hosted in

286 volcanic rocks from the Henrrieta, Mariquita, and Mesa Formations, intruded by 62.7 and

287 60.4 Ma quartz-feldspar porphyry stocks (Del Rio-Salas et al., 2013). The ore consists of a

288 nearly horizontal 60 m thick chalcocite enrichment blanket (Aponte-Barrera, 2009), which

289 is partially capped to the WSW by gravels of the Sonora Group.

290

291 2.1.4 Lucy deposit

292 The Lucy Cu-Mo deposit is located 21 km northwest of Cananea and 7 km northwest

293 of Mariquita mine (Fig. 3), accounting for resources of about 9 Mt at 0.8% Cu and 0.1% Mo

294 (Del Rio-Salas et al., 2013; González-Partida et al., 2009). The production of this deposit has

295 been sporadic, depending on the internal needs of the mining company and metal market

296 conditions. The geology in the Lucy area consists of Henrietta and Mesa Formations

297 volcanic rocks, intruded by the Cuitaca granodiorite, covered by gravel deposits of the

298 Sonora Group (Grijalva-Noriega and Roldán-Quintana, 1998).

299 The mineralization mostly occurs as breccias and disseminated chalcopyrite and

300 molybdenite, hosted within the Cuitaca granodiorite (Del Rio-Salas et al., 2013), although 301 the source of this mineralization is presently uncertain. Evidence of secondary chalcocite is

302 found as coating of disseminated pyrite. So far, no evidence of economical important

303 secondary enrichment has been reported for this deposit.

304

305 2.1.5 El Alacrán deposit

306 The El Alacrán porphyry copper prospect is located 18 km southeast of Cananea,

307 corresponding to the extreme S-SE Cananea district (Fig. 3). The El Alacrán has been

308 recognized as a low-grade deposit with resources of 0.7 Mt at 0.5% Cu (Dean, 1975). The

309 deposit is centered on a quartz-latite porphyry plug that intrudes intermediate volcanic

310 rocks of the Mesa Formation (Wodzicki, 1995).

311 Concerning the age of the mineralization, K-Ar dating on potassic alteration yielded

312 ages of 55.4 ± 1.2 (Dean, 1975) and 56.7 ± 1.2 Ma (Damon et al., 1983), which were then

313 considered as the approximation of the timing of the mineralization. However, more

314 precise data based on Re-Os molybdenite dating, yielded ages of 60.8 ± 0.2 and 60.9 ± 0.2

315 Ma in quartz-sericite veinlets (Barra et al., 2005).

316

317 2.1.6 El Pilar deposit

318 The El Pilar copper deposit is located on the southwest flank of the Sierra San

319 Antonio, which is the southern continuation of the of the Santa Cruz

320 County, Arizona (Fig. 3). The mineralization and alteration are poorly documented in

321 literature, and so far it is described as a breccia body, which is spatially related to a

322 monzonite to quartz monzonite pluton, here referred to as El Pilar pluton. The breccia

323 consists of fragments of igneous rocks sustained in a more complexly fragmented matrix of 324 the same composition, containing pyrite, chalcopyrite, and molybdenite, which was

325 partially oxidized to chrysocolla (Broch, 2012). The Sierra San Antonio is limited to the

326 west by a normal fault, which is assumed to be associated with the NNW-SSE trending

327 Basin and Range fault system, which may correspond to the limit between Sierra San

328 Antonio-Cocóspera graben. Actually, the mineralization at El Pilar is centered mostly on the

329 clastic fragments occurring as gravel, sand, and finer grains, most probably derived from

330 exhumation, partial destruction, and erosion of the mineralized breccia, due to normal

331 faulting. This is suggested by the presence of fragments of sulfide-mineralized breccia in

332 the gravels. According to Broch (2012), the thickness of productive gravel ranges from 30

333 to 180 m. The El Pilar resource accounts for 0.36 Mt at 0.15% Cu.

334

335 3. ANALYTICAL PROCEDURES

336 3.1. Re-Os method

337 The Re–Os analyses were conducted following the method described in Mathur et al.

338 (2002) and Teixeira-Correia et al. (2007). Approximately 0.05–0.1 g of handpicked

339 molybdenite was loaded in a Carius tube with 10 ml of reverse aqua regia. While the

340 reagents, sample and spikes were frozen, the Carius tube was sealed and left to thaw at

341 room temperature. The tube was placed in an oven and heated to 240°C overnight.

342 Once the Carius tube was opened after the high temperature equilibration, 4 mL of

343 CCl4 solvent is added on top of the acid solution while still frozen. Once thawed, the aqua

344 regia solvent mixture is transferred to 50 mL falcon tubes, agitated and centrifuged to aid

345 the extraction and separation of Os into organic-solvent and acidic layers. The organic-

346 solvent layer containing the Os is separated from the acidic layer and the procedure is 347 repeated twice more with 3 mL of CCl4 to ensure high Os yield. The Os is then back

348 extracted from the organic-solvent into concentrated hydrobromic acid, which is dried and

349 subsequently purified for mass spectrometry by microdistillation techniques (Birk et al.,

350 1997). Osmium was further loaded on platinum filaments with Ba(OH)2 to enhance

351 ionization. After osmium separation, the remaining acid solution was dried and later

352 dissolved in 0.1 HNO3. Rhenium was extracted and purified through a two-stage column

353 using AG1-X8 (100–200 mesh) resin and loaded on platinum filaments with BaSO4.

354 Samples were analyzed by negative thermal ion mass spectrometry (N-TIMS)

355 (Creaser et al., 1991) on a VG 54 mass spectrometer. Molybdenite ages were calculated

356 using a 187Re decay constant of 1.666x10-11 per year (Smoliar et al., 1996). Ages are

357 reported with a 0.5% error, which is considered a conservative estimate and reflects all

358 sources of error (i.e. uncertainty in the Re decay constant (0.31%), 185Re and 190Os spike

359 calibrations (0.08% and 0.15%, respectively), analytical and weighing errors.

360

361 3.2. U-Pb method

362 The determination of the ages in zircons was analyzed following the procedure

363 described in Del Rio-Salas et al. (2013), and the description is outlined below. Around 1 kg

364 of the intrusive rocks were crushed and milled. Heavy mineral concentrates smaller than

365 350 µm were separated using the Wilfley Table. The zircons were concentrated using di-

366 iodomethane heavy liquid and magnetic techniques. Later, the zircons were handpicked

367 under a binocular microscope, and mounted in an epoxy resin and polished. Around 30

368 zircons from each sample were analyzed by laser ablation multicollector inductively

369 coupled plasma mass spectrometry (LA-MC-ICPMS) at the Arizona LaserChron Center. The 370 analyses involve ablation of zircon with a New Wave/Lambda Physik DUV193 Excimer

371 laser (operating at a wavelength of 193 nm) using a spot diameter of 35 microns. The

372 ablated material is carried with helium gas into the plasma source of a GV Instruments

373 Isoprobe, which is equipped with a flight tube of sufficient width for simultaneous

374 measurements of U, Th, and Pb isotopes. All measurements are made in static mode, using

375 Faraday detectors for 238U and 232Th, an ion-counting channel for 204Pb, and either faraday

376 collectors or ion counting channels for 208-206Pb. Ion yields are ~1 mv per ppm. Each

377 analysis consists of one 20-second integration on peaks with the laser off (for

378 backgrounds), 20 one-second integrations with the laser firing, and a 30-second delay to

379 purge the previous sample and prepare for the next analysis. The ablation pit is ~15

380 microns in depth.

381 For each analysis, the errors in determining 206Pb/238U and 206Pb/204Pb result in a

382 measurement error of ~1% (at 2-sigma level) in the 206Pb/238U age. The errors in

383 measurement of 206Pb/207Pb and 206Pb/204Pb also result in ~1% (2-sigma) uncertainty in

384 age for grains that are >1.0 Ga, but are substantially larger for younger grains because of

385 the low intensity of the 207Pb signal. For most analyses, the cross-over in precision of

386 206Pb/238U and 206Pb/207Pb ages occurs at ~1.0 Ga.

387 Common Pb correction is accomplished by using the measured 204Pb and assuming

388 an initial Pb composition from Stacey and Kramers (1975) (with uncertainties of 1.0 for

389 206Pb/204Pb and 0.3 for 207Pb/204Pb). Our measurement of 204Pb is unaffected by the

390 presence of 204Hg because backgrounds are measured on peaks (thereby subtracting any

391 background 204Hg and 204Pb), and because very little Hg is present in the argon gas. 392 Inter-element fractionation of Pb/U is generally ~20%, whereas fractionation of Pb

393 isotopes is generally <2%. In-run analysis of fragments of a large zircon crystal (generally

394 every fifth measurement) with known age of 564 ± 4 Ma (2-sigma error) is used to correct

395 for this fractionation. The uncertainty resulting from the calibration correction is generally

396 ~1% (2-sigma) for both 206Pb/207Pb and 206Pb/238U ages. The analytical data are reported

397 in Tables 3-9. Uncertainties shown in these tables are at the 1-sigma level, and include only

398 measurement errors.

399 The reported ages are determined from the weighted mean (Ludwig, 2003) of the

400 206Pb/238U or 206Pb/207Pb ages of the concordant and overlapping analyses. Analyses that

401 are statistically excluded from the main cluster are shown in blue on these figures. Two

402 uncertainties are reported on these plots. The smaller uncertainty (labeled “mean”) is

403 based on the scatter and precision of the set of 206Pb/238U or 206Pb/207Pb ages, weighted

404 according to their measurement errors (shown at 1-sigma). The larger uncertainty (labeled

405 “age”), which is the reported uncertainty of the age, is determined as the quadratic sum of

406 the weighted mean error plus the total systematic error for the set of analyses. The

407 systematic error, which includes contributions from the standard calibration, age of the

408 calibration standard, composition of common Pb, and U decay constants, is generally ~1-

409 2% (2-sigma).

410

411 4. RESULTS

412 4.1. Re-Os geochronological data

413 The new Re-Os molybdenite age for the mineralization from Mariquita, Lucy, and El

414 Pilar deposits are reported in Table 2. A molybdenite sample from the El Pilar deposit 415 yielded a total rhenium and 187Os concentrations of 64.8 ppm and 50.2 ppb, respectively.

416 The data yields a molybdenite age of 73.9 ± 0.3 Ma, which is the oldest age determined in

417 this study, and so far, the oldest mineralization age reported within the Cananea district.

418 The total rhenium and 187Os concentrations for two molybdenite samples from the

419 Mariquita deposit range from 83.7 to 373.5 ppm and 51.6 to 231.6 ppb, respectively. The

420 Re-Os isotope data yield a molybdenite mineralization age between 59.2 and 59.3 ± 0.3 Ma

421 for the molybdenite mineralization at Mariquita deposit.

422 The total rhenium and 187Os concentrations for a molybdenite sample from the Lucy

423 deposit are 47.2 ppm and 29.7 ppb, respectively. The Re-Os data yield an age of 61.8 ± 0.3

424 Ma for the mineralization at Lucy deposit.

425

426 4.2. U-Pb zircon data

427 The U-Pb zircon ages are shown in Tables 3-9. All reported ages have uncertainties

428 at the two-sigma level, which only includes the analytical error. The age of each sample

429 includes additional uncertainties from the calibration correction, decay constant and

430 common lead. These systematic errors (<1.4 %) are added quadratically to the analytical

431 error. The analyzed zircons from the intrusive rocks from the Cananea district have U

432 concentrations that range from 2300-180 ppm. All zircons yield U/Th ratios of ~2,

433 characteristic of igneous zircons (Rubatto, 2002).

434 Tables 3-6 show the U-Pb zircon ages for the Buenavista del Cobre porphyry

435 samples. Table 3 shows the U-Pb zircon ages for a granodiorite porphyry sample from

436 Buenavista del Cobre, which yielded a weighted average age of 60.8 ± 1.0 Ma (n = 22,

437 MSWD 3.8; Fig. 4a). Zircons from this sample produced an inherited ages from Middle 438 Jurassic (167 Ma, n = 1), Late Cretaceous (73 Ma, n = 1), and Early Paleocene (~64 Ma, n =

439 3). Zircons from a granodiorite porphyry sample yielded a weighted average age of 60.9 ±

440 1.2 Ma (n = 18, MSWD 1.4; Fig. 4b, Table 4). The U-Pb zircon data show inherited ages of

441 Early Jurassic (190 Ma, n = 2), Middle Jurassic (165 Ma, n = 3), Late Cretaceous (68 Ma, n =

442 2), Early Paleocene (64 Ma, n = 3).

443 The zircons from quartz monzonite porphyry sample yielded a weighted average

444 206Pb/238U age of 61.3 ± 1.4 Ma (n=16, MSWD =2.4; Fig. 4c, Table 5). Zircons from this

445 sample produced inherited ages from Early Jurassic (195 Ma, n = 1), Middle Jurassic (170

446 Ma, n = 1), Early Cretaceous (140 and 124 Ma), Late Cretaceous (74 Ma, n = 1), and Early

447 Paleocene (~64 Ma, n = 6). A sample from younger monzodiorite porphyry in Buenavista

448 del Cobre yielded a weighted average age of 58.9 ± 1.4 Ma (n = 24, MSWD 4.5; Fig. 4d, Table

449 6). Zircons from this porphyritic sample yielded inherited ages from Late Ordovician (458

450 Ma, n = 1), Early Devonian (396 Ma, n = 1), Early Cretaceous (100 Ma, n = 1), and Late

451 Cretaceous (90 Ma, n = 1).

452 A quartz-monzonitic porphyry sample from the El Alacrán deposit yielded an age of

453 57.8 ± 1.0 Ma (n = 14, MSWD 1.8; Fig. 5a, Table 7). Zircons from this sample yielded

454 inherited ages from Early Cretaceous (124 Ma, n = 1), Late Cretaceous (68 Ma, n = 1), Early

455 Paleocene (~64 Ma, n = 5).

456 Two U-Pb zircon ages from the granodiorite from the El Pilar deposit are shown in

457 Tables 8 and 9. U-Pb zircon data yielded ages of 74.6 ± 1.4 Ma (n = 27, MSWD 3; Fig. 5b,

458 Table 8) and 74.7 ± 1.1 Ma (n = 31, MSWD 2.9; Fig. 5c, Table 9). Only one inherited zircon

459 age from Early Cretaceous (116 Ma) was found.

460 461 5. DISCUSSION

462 5.1 U-Pb and Re-Os geochronology

463 There is a relatively large geochronological dataset of the Cananea mining district

464 (e.g. Damon and Mauger, 1966; Damon et al., 1983), however uncertainties due to

465 differences in closure temperatures of dated materials, analytical procedures and precision,

466 etc., precludes straightforward interpretations of the magmatic and hydrothermal activity.

467 Thus, filtering of the data is required in order to reproduced consistent evolutionary

468 models. The new ages reported in this study, together with other available U-Pb zircon and

469 Re-Os molybdenite dates (Barra et al., 2005; Noguez-Alcántara, 2008; Valencia et al., 2006)

470 allow to constrain more plausible interpretation regarding ages and relationships between

471 magma crystallization and mineralizing hydrothermal pulses.

472 Our results indicate that magmatic activity in the Cananea mining district occurred

473 in a time span between 75 and 58 Ma. The oldest age corresponds to a monzonite to quartz

474 monzonite pluton exposed in the area of El Pilar, in the northwestern part of the district,

475 which yielded ages of 74.6 ± 1.4 and 74.7 ± 1.1 Ma (Fig. 5b,c). These ages are in good

476 agreement with the U-Pb zircon 74.0 ± 1.1 Ma old Washington camp stock, exposed in the

477 Patagonia Mountains (Vikre et al., 2014), ~20 km north of El Pilar area. Also, the El Pilar

478 deposit yielded the oldest age for the Cu-Mo mineralization within the district, with a Re-Os

479 molybdenite date of 73.9 ± 0.3 Ma. These very similar ages indicate that the mineralizing

480 pulse was nearly synchronic with the emplacement of the El Pilar pluton, suggesting a

481 cogenetic relationship.

482 In the central part of the district, the 64 Ma old Cuitaca granodiorite (Anderson and

483 Silver, 1977) is recalled as the precursor of the Laramide mineralizing porphyritic 484 intrusions (Noguez-Alcántara, 2008; Wodzicki, 1995). This igneous body is exposed over

485 an area exceeding 400 km2 (Fig. 3), forming the bulk of the Sierras El Chivato, Cuitaca, and

486 the western flank of the Sierra Mariquita and Sierra Elenita.

487 The oldest mineralizing porphyritic intrusion recognized within the district

488 corresponds to a quartz monzonite plug that controlled the mineralization in the Milpillas

489 porphyry copper deposit (PCD). U-Pb zircon data of sample from this plug yielded an age of

490 63.9 ± 1.3 Ma (Valencia et al., 2006), which is quite similar to the age of 64 Ma reported for

491 the Cuitaca granodiorite (Anderson and Silver, 1977). The hypogene mineralization in the

492 Milpillas PCD is constrained by two Re-Os molybdenite ages of 63.0 ± 0.4 and 63.1 ± 0.4 Ma,

493 suggesting a clear temporal correlation between the hydrothermal activity and the

494 porphyritic mineralizing intrusion (Valencia et al., 2006).

495 In the case of the Lucy deposit, a sample of the Cuitaca granodiorite collected in the

496 Lucy pit, which is the host of the disseminated Cu-Mo mineralization, yielded an U-Pb

497 zircon age of 63.8 ± 1.1 Ma (Del Rio-Salas et al., 2013). The mineralizing event is

498 constrained by a Re-Os molybdenite age of 61.8 ± 0.3 Ma (Table 2), which reveals a wider

499 time span between the plutonic phase and the mineralization in Lucy (~2 Ma).

500 In the Mariquita area, the Cuitaca granodiorite is widespread, unfortunately this

501 rock has not been dated yet for this place, however, the age of 64 Ma reported by Anderson

502 and Silver (1977), comes from a sample collected 5 km southeast of Mariquita mine. So far,

503 the mineralization appears to be related to intrusions of two porphyritic stocks, which

504 yielded U-Pb zircon ages of 62.7 ± 1.3 and 60.4 ± 1.1 Ma (Del Rio-Salas et al., 2013). Here,

505 the mineralization is constrained by a Re-Os age of 59.3 ± 0.3 Ma (Table 2), from a

506 molybdenite separates collected from the younger porphyritic intrusion. 507 In the Buenavista del Cobre PCD, in the southeastern part of the district, the Cuitaca

508 granodiorite is not recognized, however it outcrops ~3 km to the west-southwest of the

509 mining area (Fig. 3). This deposit is not only the largest within the Cananea district, but also

510 the most complex in terms of the magmatic and hydrothermal histories. Here, at least four

511 lithologically distinctive porphyritic phases have been recognized and dated. U-Pb zircon

512 data in three of them yielded fairly similar ages between 60.8 ± 1.0 to 61.3 ± 1.4 Ma. The

513 fourth one yielded a younger age of 58.9 ± 1.4 Ma (Tables 3–6). The age of the

514 mineralization has been constrained by Re-Os molybdenite dates of 59.3 ± 0.3 and 59.2 ±

515 0.3 Ma (Barra et al., 2005). Field evidences do not offer definitive clues regarding the

516 contribution to the metal budget.

517 The El Alacrán represents the southernmost PCD within the Cananea district. Here,

518 the Cuitaca granodiorite is not exposed, but two quartz-monzonitic porphyries (Arellano-

519 Morales, 2004) are found intruding andesitic to quartz latite volcaniclastic rocks (Dean,

520 1975) correlatable with the Mesa Formation (Cox et al., 2006). A drill core sample from one

521 of the mineralized porphyritic intrusion yielded an U-Pb zircon age of 57.8 ± 1.0 Ma (Table

522 7). This age is younger than previously reported Re-Os molybdenite ages of 60.9 ± 0.2 and

523 60.8 ± 0.2 Ma (Barra et al., 2005). These Re-Os ages are inconsistent with the age of the

524 dated plug, suggesting that the mineralization corresponds to a different magmatic event,

525 perhaps correlatable with the other, undated, porphyritic plug.

526

527 5.2 Southeastern progression of Cu-Mo mineralization

528 The distribution of most of the Cordilleran porphyry copper deposits in

529 southwestern North America lie along a ~350 km northwest-trending regional lineament, 530 extending from Silver Bell mine, Arizona, to La Caridad mine, Sonora, which was formerly

531 referred to as the Cananea lineament (Hollister, 1978). The Mexican part of this lineament

532 includes the Cananea and Nacozari mining districts, extending for ~140 km from El Pilar

533 deposit to La Caridad mine.

534 In the Cordilleran porphyry copper belt, two main episodes of mineralization have

535 been previously documented at 74-70 and 60-55 Ma (McCandless et al., 1993). As

536 mentioned above, these episodes are also reported for the Cananea and Nacozari mining

537 districts, where ages range from 74 to 54 Ma. However, the most significant number of ages

538 lies in the range of 60-55 Ma (Barra et al., 2005).

539 Interestingly enough, a southeastward decrease of molybdenite mineralization ages

540 is discernible along the Cananea lineament, from the El Pilar deposit (~74 Ma), Milpillas

541 (~63 Ma), Lucy (~62 Ma), Buenavista del Cobre and Mariquita (~59 Ma), to La Caridad

542 deposit (~54 Ma). The only deposit that apparently does not follow this pattern is El

543 Alacrán, which yielded an age ~61 Ma. However, the slight difference in age may be partly

544 enhanced by analytical uncertainties. This southeastward decrease in the mineralization

545 ages in some way reflects a similar decrease in the emplacement ages of the mineralizing

546 intrusions (Fig. 6). The relatively minor age difference in the porphyritic intrusions and the

547 primary sulfide ores suggests a NW-SE continuum in the magmatic and hydrothermal

548 processes along the Cananea lineament.

549 The distribution of mineralization ages in the Cananea and Nacozari mining districts

550 can be coupled to the well-accepted east-northeastward migration model of the Laramide

551 magmatic arc (Coney and Reynolds, 1977; Damon et al., 1983). However, there is no doubt 552 that a southeastern migration in the Cu-Mo mineralization occurred within these districts,

553 which gives support to the Cananea lineament (Fig. 6).

554 The U-Pb zircon age of the El Pilar monzonitic pluton (~74 Ma), along with previous

555 U-Pb zircon ages, suggest a time span of ~17 Ma for the magmatic activity in the Cananea

556 district. On the other hand, the available U-Pb zircon ages suggest that the age of

557 emplacement of the porphyritic stocks occurred within a shorter period of time of about 6

558 Ma. Moreover, the Re-Os molybdenite ages allow identification of at least five discrete

559 mineralizing pulses, which took place within ~4 Ma (Fig. 6).

560

561 5.3 Tertiary tectonics

562 In the Cananea district as well as most northeastern Sonora, Mesozoic volcanoclastic

563 rocks and Laramide intrusions lack deep extensional structures, such as cataclastic fault

564 zones or low angle mylonitic detachment faults, which suggests the amount of extension

565 was low to moderate during the Basin and Range extensional tectonics. This supports the

566 conclusions reached by Nourse et al. (1994), regarding the NW-SE trending Imuris

567 lineament, which separates the Cenozoic Metamorphic Core Complex belt to the southwest,

568 from the less-deformed and unmetamorphosed region to the northeast, which

569 encompasses the Cananea mining district (Fig. 1).

570 Another argument supporting a low to moderate extension is the presence of

571 shallow porphyry copper deposits in both horsts and grabens. If a horst of the Cananea

572 mining district were limited by a low-angle detachment fault, rocks of the middle and lower

573 crust would be expected to be exposed, meaning that the shallower levels of the crust,

574 including the porphyry copper systems, should have been removed. The 1.4 Ga Cananea 575 granite in the Mariquita, El Cobre, and Elenita ranges, as well as the Pinal schist outcrops in

576 the Sierra Los Ajos, are interpreted as the result of uplift during the Laramide

577 compressional phase along high-angle reverse thrust faults, and not as a consequence of

578 tectonic denudation of the upper crust along a low-angle normal fault. The main direction

579 of extension is ~NE60°SW, which is highlighted by similarly oriented transfer faults

580 reported in Arizona (Wilkins Jr. and Heidrick, 1995), and Sonora (Calmus et al., 2011; Gans,

581 1995). Moreover, this structural pattern is also underlined by direction of metamorphic

582 lineation and S-C shear zones observed in the metamorphic core complexes mylonitic

583 zones (Vega-Granillo and Calmus, 2003; Wilkins Jr. and Heidrick, 1995). In the Cananea

584 mining district this structural pattern is probably exemplified by the lineament between

585 Imuris and Milpillas, which may represent an accommodation zone during the Basin and

586 Range extension (Fig. 7). In the Milpillas-Cuitaca graben, Carreón-Pallares (2002)

587 considered that the Milpillas and Mariquita areas formed two subdistricts, which were

588 dissected during two main Miocene to Pleistocene extensional events. The first one is

589 characterized by N30°E left-lateral faults that segmented these subdistricts, whereas the

590 second one dealt with normal faulting associated with formation of the Milpillas-Cuitaca

591 graben. According to this model, the last event caused an ~5 km offset between Lucy and

592 Milpillas mines, which is questionable if only dip-slip displacements along high-angle

593 normal faults are considered.

594

595 5.4 Implications for regional exploration

596 Traditionally, the importance of PCD lies on the presence of supergene-enriched

597 blankets, due to the leaching of primary sulfide mineralization by meteoric water (Chávez, 598 2000). In the Cananea mining district, this process was favored by the Basin and Range

599 extensional tectonics, enhanced by climatic conditions, commonly displaying multiple

600 copper enrichment cycles. Further sedimentary processes stand as critical factors to

601 preserve the secondary sulfide ores. In northwestern Mexico, examples displaying different

602 levels of erosion support this assumption. For instance, the Buenavista del Cobre mine was

603 apparently very well preserved, so most of the system is present. On the contrary, there are

604 examples where the system is more deeply eroded, such as in the case of El Crestón Mo-Cu

605 deposit in central Sonora (Valenzuela-Navarro et al., 2005). Some of the PCD within the

606 Cananea district are partially or fully covered by basin-fill clastic sediments of the Báucarit

607 Formation and the Sonora Group (Grijalva-Noriega and Roldán-Quintana, 2000), such as in

608 the Milpillas mine, where copper ores are being mined underground.

609 Cenozoic block faulting tectonics was important in the preservation vs. erosion

610 processes in PCD from northwestern Mexico and southeastern North America (Barton et

611 al., 1995). The extensional tectonics in Sonora, and particularly in the Cananea mining

612 district, may have dissected important parts of the porphyry copper systems, and even

613 more, it may have buried some systems completely, so that hidden, and potentially

614 important deposits, may remain undiscovered, becoming important targets for future

615 mineral exploration along the Cananea lineament, from El Pilar deposit at the northwest,

616 through the La Caridad deposit, at the southeastern section of the lineament.

617 In some cases, the PCD underwent geological processes where copper-bearing acid

618 solutions flow laterally along paleochannels in gravel deposits (Chávez, 2000). Copper-rich

619 solutions interact with gravel precipitating different copper mineral species controlled by

620 pH changes, currently referred to as exotic copper deposits, such as La Exótica, which is an 621 important deposit adjacent to the giant Chuquicamata PCD in northern Chile (Münchmeyer,

622 1996). Even though, no deposits of this type have been recognized in the Cananea district,

623 there are some evidences, particularly in the Milpillas and the Mariquita deposits, where

624 secondary copper minerals, including copper carbonates, silicates, and oxides, suggest that

625 formation of exotic copper may be a possibility along the Cananea lineament.

626

627 6. CONCLUSIONS

628 The Re-Os molybdenite ages from the Cananea district suggest at least five well-

629 constrained mineralizing pulses at ~74, 63, 62, 60, and 59 Ma, nevertheless, the main

630 mineralization stage occurred in a much shorter period of time of ~4 Ma. Also, the new Re-

631 Os molybdenite age for the El Pilar deposit (~74 Ma) records the oldest mineralizing pulse

632 reported so far in the Cananea district, establishing the time for the initiation of the

633 Laramide porphyry copper mineralization in northern Sonora.

634 The U-Pb zircon ages of the mineralizing porphyries suggest a continued magmatic

635 period of ~6 Ma. Including the new U-Pb zircon ages of the hosting rock from the El Pilar

636 deposit, the magmatic activity expands to ~17 Ma. Similar mineralizing and magmatic ages

637 reported in the Patagonia Mountains, in southern Arizona (Vikre et al., 2014), suggest that

638 this magmatic-hydrothermal activity occurred synchronically in northern Sonora and

639 southern Arizona.

640 Interestingly enough, the new ages indicate a NW-SE progression, from older to

641 younger ages, in both, the mineralizing magmatic intrusions and the mineralization at least

642 from El Pilar to La Caridad (Fig. 2), suggesting the existence of a regional structural control

643 of the mineralization emplacement along the Cananea lineament. This is an important 644 contribution that may bring new clues to understand the dynamics of formation and

645 regional emplacement of the PCD.

646 Further structural studies to better elucidate the Cenozoic extensional tectonics that

647 dissected and rearranged the deposits, may help not only to reconstruct the original

648 systems, but also provide hints to enhance exploration for hidden PCD and exotic copper

649 mineralization in the Basin and Range province.

650

651 7. ACKNOWLEDGMENTS

652 This research was supported by the CONACYT (project 166600), the consortium between

653 FRISCO and the Geology Department at the University of Sonora, and the Geosciences

654 Department at the University of Arizona. We thank Arizona LaserChron Center, especially

655 George Gehrels for the support. We thank Fernando Barra for preliminary Re-Os

656 molybdenite analysis. We are grateful to Grupo FRISCO and Maria Mine staff for the

657 logistics and support. We thank Ramon Ayala from Grupo México for sampling support in

658 the Cananea Mine. This work has been partly funded by the Universidad Nacional Autónoma de

659 México trough the DGAPA Program, which supported the sabbatical year of TC between July

660 2011 and July 2012 at the Thermochronology Laboratory of ISTerre, University Joseph Fourier,

661 Grenoble), during which studies were conducted on the Cenozoic exhumation of porphyry

662 copper deposits in Sonora. We are thankful for the field assistance of the geologists Roman

663 Solís, Oscar Saitz Sau, Cruz Páez, and Julio Cesar Orantes. We thank the editor(s) and the

664 referees.

665

666 667

668 8. REFERENCES

669 Amato, J.M., Boullion, A.O., Serna, A.M., Sanders, A.E., Farmer, G.L., Gehrels, G.E., Wooden,

670 J.L., 2008. Evolution of the Mazatzal province and the timing of the Mazatzal orogeny:

671 Insights from U-Pb geochronology and geochemistry of igneous and metasedimentary

672 rocks in southern New Mexico: Geol. Soc. Amer. Bull. 120, 328–346.

673

674 Anderson, T.H., Silver, T.H., 1977. U-Pb isotope ages of granitic plutons near Cananea,

675 Sonora. Econ. Geol. 72, 827–36.

676

677 Anderson, T. H., Silver, T. H., 1979. The role of the Mojave-Sonora megashear in the tectonic

678 evolution of northern Sonora. Geology of northern Sonora. Geol. Soc. Ame. Field Trip

679 Guidebook, 7, 59-68.

680

681 Aponte-Barrera, M., 2009. Geología y mineralización del yacimiento Mariquita, distrito de

682 Cananea, in Clark, K.F., Salas-Pizá, G.A., Cubillas-Estrada, R. (Eds.), Geología Económica de

683 México. Servicio Geológico Mexicano, pp. 852–856.

684

685 Arellano-Morales, R., 2004. Caracterización Geoquímica y estudio de inclusiones fluidas del

686 prospecto El Alacrán, Cananea, Sonora, México. Unpublished M.S. Thesis, Universidad de

687 Sonora, Hermosillo Sonora, 107 p.

688 689 Ayala-Fontes, 2009. Geología y mineralización en el distrito minero Cananea, Sonora,

690 México, in Clark, K.F., Salas-Pizá, G.A., Cubillas-Estrada, R. (Eds.), Geología Económica de

691 México. Servicio Geológico Mexicano, p. 804–810.

692

693 Barra, F., Ruiz, J., Valencia, V.A., Ochoa-Landín, L., Chesley, J., 2005. Laramide Porphyry Cu-

694 Mo Mineralization in Northern Mexico: Age Constraints from Re-Os Geochronology in

695 Molybdenite. Econ. Geol. 100, 1605–16.

696

697 Barton, M.D., Staude, J.M., Zürcher, L., Megaw, P.K.M., 1995. Porphyry copper and other

698 intrusion-related mineralization in Mexico, in Pierce, F.W. Bolm, J.G., (Eds.), Porphyry

699 copper deposits of the American Cordillera. Arizona Geological Society Digest. 20, 487–524.

700

701 Birck, J.L., Barman, M.R., Capmas, F., 1997. Re‐Os Isotopic Measurements at the Femtomole

702 Level in Natural Samples. Geostandards newsletter. 21, 19–27.

703

704 Broch, M., 2012. El Pilar Project, N143-101F1. Techbical Report, feasibility study, Mercator

705 Minerals Ltd. 302 p.

706

707 Busby-Spera, C.J., 1988. Speculative tectonic model for the early Mesozoic arc of the south-

708 west Cordilleran United States. Geology. 16, 1121–1125.

709

710 Bushnell, S.E., 1988. Mineralization at Cananea, Sonora, Mexico, and the paragenesis and

711 zoning of breccia pipes in Quartz-feldspathic rock. Econ. Geol. 83, 1760–81. 712

713 Calmus, T., Vega-Granillo, R., Lugo-Zazueta, R., 2011. Evolución geológica de Sonora durante

714 el Cretácico Tardío y el Cenozoico, in Calmus, T., (Ed.), Panorama de la Geología de Sonora,

715 México. Universidad Nacional Autónoma de México, Instituto de Geología, Boletín 118,

716 227–266.

717

718 Carreón-Pallares, J.N., 2002. Structure and tectonic history of the Milpillas porphyry copper

719 district, Sonora, Mexico. Unpublished M.S. Thesis, University of Utah, Salt Lake City, 72 p.

720

721 Chávez, W.X., 2000. Supergene oxidation of copper deposits: zoning and distribution of

722 copper oxide minerals. Soc. Econ. Geol. Newsletter. 41, 1–21.

723

724 Creaser, R.A., Papanastassiou, D.A., Wasserburg, G.J., 1991. Negative thermal ion mass

725 spectrometry of osmium, rhenium and iridium. Geochim. Cosmochim. Acta. 55, 397–401.

726

727 Coney, P.J., Reynolds, S.J., 1977. Cordilleran benioff zones. Nature. 270, 403–406.

728

729 Cox, D.P., Miller, R.J., Woodbourne, K.L., 2006. The Laramide Mesa Formation and the Ojo de

730 Agua Caldera, Southeast of the Cananea Copper Mining District, Sonora, Mexico. U. S.

731 Geological Survey.

732

733 Damon, P.E., Mauger, R.L., 1966. Epeirogeny-orogeny viewed from the Basin and Range

734 Province. Aime Trans. 235, 99–112. 735

736

737 Damon, P.E., Shafiqullah, M., Clark, K.F., 1983. Geochronology of the porphyry copper

738 deposits and related mineralization of Mexico. Can. J. Earth Sci., 20, 1052–71.

739

740 Davis, G.H., 1979. Laramide folding and faulting in southeastern Arizona. Am. J. Sci. 279,

741 543–569.

742

743 Dean, D.A., 1975. Geology, alteration, and mineralization of the El Alacran area, Northern

744 Sonora, Mexico. Unpublished M.S. Thesis, University of Arizona, Tucson Arizona, 222 pp.

745

746 Del Rio-Salas, R., Ochoa-Landín, L, Ruiz, J., Eastoe, C., Meza-Figueroa, D., Zuñiga-Hernández,

747 H., Mendívil-Quijada, H., Quintanar-Ruiz, F., 2013. Geology, stable isotope, and U-Pb

748 geochronology of the Mariquita porphyry copper and Lucy Cu-Mo deposits, Cananea

749 District, Mexico: A contribution to regional Exploration. J. Geochem. Explor. 124, 140–154.

750

751 Einaudi, M.T., 1982. Description of skarns associated with porphyry copper plutons, south–

752 western North America, in Titley, S.R. (Ed.), Advances in Geology of the Porphyry Copper

753 Deposits, South–western North America. The University of Arizona Press, 139–175.

754

755 Gans, P.B., 1997. Large-magnitude Oligo-Miocene extension in southern Sonora—

756 Implications for the tectonic evolution of northwest Mexico. Tectonics. 16, 388–408.

757 758 Grajales-Nishimura, J.M., Terrell, D.J., Damon, P.E., 1992. Evidencias de la prolongacion del

759 arco magmatico Cordillerano del Triasico Tardio-Jurásico en , Durango y

760 Coahuila. Boletín de la Asociación Mexicana de Geólogos Petroleros. 42, 1–18.

761

762 Grijalva-Noriega, F.J., Roldán-Quintana, J., 1998. An overview of the Cenozoic tectonic and

763 magmatic evolution of Sonora, northwestern Mexico. Rev. Mex. Ciencias Geológicas. 15,

764 145–156.

765

766 González-Partida, E., Pérez-Segura, E., Camprubi, A., Lhomme, T., 2009. Ore-forming fluids

767 in the Lucy porphyry Cu–Mo deposit and its regional significance in the Cananea district

768 (Sonora, Mexico). J. Geochem. Explor. 101, p. 39.

769

770 González-Partida, E., Pérez-Segura, E., Camprubi, A., González-Ruíz, L., 2013.

771 Espectrometría Raman y microtermometría de inclusiones fluidas en cuarzo magmático e

772 hidrotermal en los pórfidos de Cu-Mo de Lucy y María, Sonora, México. Rev. Mex. Ciencias

773 Geológicas, 30, 198–209.

774

775 Henry, C.D., McDowell, F.W., Silver, L.T., 2003. Geology and geochronology of granitic

776 batholithic complex, Sinaloa, Mexico—implications for Cordilleran magmatism and

777 tectonics, in Johnson, S.E., Paterson, S.R., Fletcher, J.M., Girty, G.H., Kimbrough, D.L., Martín-

778 Barajas, A., (Eds.), Tectonic evolution of northwestern Mexico and the southwestern USA.

779 Geol. Soc. Spec. Pap. 374, 237–273.

780 781 Hollister, V.F., 1978. Geology of the porphyry copper deposits of the western Hemisphere.

782 New York, Soc. Mining Engineers AIME, 219 p.

783

784 Keith, S.B. Swan, M.M, 1996. The great Laramide porphyry copper cluster of Arizona,

785 Sonora, and New Mexico: The tectonic setting, petrology and genesis of the world class

786 metal cluster, in: Coyner, A.R., Fahey, P.L. (Eds.), Geology and ore deposits of the American

787 cordillera: Geological Society of Nevada Symposium Proceedings, Reno-Sparks, Nevada, pp.

788 1667–1747.

789

790 Ludwig, K.R., 2003, Isoplot 3.00. Berkeley Geochronology Center, Special Publication 4, 70

791 p.

792

793 Mathur, R., Marschik, R., Ruiz, J., Munizaga, F., Leveille, R., 2002. Age of mineralization of the

794 Candelaria iron oxide Cu–Au deposit, and the origin of the Chilean Iron Belt based on Re–Os

795 isotopes. Econ. Geol. 97, 59–71.

796

797 Meinert, L.D., 1982. Skarn, Manto, and Breccia Pipe Formation in Sedimentary Rocks of the

798 Cananea Mining District, Sonora, Mexico. Econ. Geol. 77, 919–9.

799

800 McCandless, T.E., Ruiz, J., Campbell, A.R., 1993. Rhenium behavior in molybdenite in

801 hypogene and near-surface environments: Implications for Re-Os geochronology: Geochim.

802 Cosmochim. Acta. 57, 889–905.

803 804 Mulchay, R.B., Velasco, J.R., 1954. Sedimentary rocks at Cananea, Sonora, Mexico and

805 tentative correlation with the sections at Bisbee and the , Arizona.

806 Mining Eng. 6, 628–632.

807

808 Münchmeyer, C., 1996. Exotic deposits-products of lateral migration of supergene solutions

809 from porphyry copper deposits, in Camus, F., Sillitoe, R.H., Petersen, R. (Eds.), Andean

810 copper deposits: new discoveries, mineralization, styles and metallogeny. Soc. Econ. Geol.

811 Spec. Pub. 5, 43–58.

812

813 Noguez-Alcántara, B., 2008. Reconstrucción del modelo genético y evolución tectónica del

814 yacimiento tipo pórfido cuprífero Milpillas, Distrito de Cananea, Sonora, México.

815 Unpublished Ph.D. Thesis, Universidad Nacional Autónoma de México, 390 p.

816

817 Noguez-Alcántara, B., Valencia-Moreno, M., Roldán-Quintana, J., Calmus, T., 2007.

818 Enriquecimiento supergénico y análisis de balance de masa en el yacimiento de pórfido

819 cuprífero Milpillas, Distrito Cananea, Sonora, México. Rev. Mex. Ciencias Geologicas. 24,

820 368–388.

821

822 Nourse, J.A., Anderson, T. H., Silver, L. T., 1994. Tertiary metamorphic core complexes in

823 Sonora, northwestern Mexico. Tectonics. 13, 1161–1182.

824

825 Ochoa-Landín, L. and Echavarri, A., 1978. Observaciones preliminares sobre la

826 secuencia de las intrusiones hipabisales en el tajo Colorado-Veta del Distrito 827 Minero de Cananea: Boletín del Departamento de Geología de la Universidad de

828 Sonora, 1, 57–60.

829

830 Ortega-Rivera, A., 2003, Geochronological constraints on the tectonic history of the

831 Peninsular Ranges Batholith of Alta and Baja California—tectonic implications for western

832 Mex-ico, in Johnson, S. E., Paterson, S.R., Fletcher, J.M., Girty, G.H., Kimbrough, D.L., Martín-

833 Barajas, A., (Eds.), Tectonic evolution of northwestern Mexico and the southwestern USA.

834 Geol. Soc. Spec. Pap. 374, 297–335.

835

836 Ochoa-Landín, L., Echávarri, A., 1978. Observaciones preliminares sobre la secuencia de las

837 intrusiones hipabisales en el tajo Colorada-Veta del distrito minero de Cananea. Boletín del

838 Departamento de Geología de la Universidad de Sonora. 1, 57–60.

839

840 Pérez-Segura E., 1985. Carta Metalogenética de Sonora 1:250,000—una interpretación de

841 la metalogenia de Sonora. Gobierno del Estado de Sonora Publicación 7, 1985, 64 p.

842

843 Rodríguez-Castañeda, J.L., Anderson, T.H., 2011. El arco magmático jurásico en Sonora,

844 México—Distribución, edades y ambiente tectónico, in Calmus, T., (Ed.), Panorama de la

845 geología de Sonora, México. Universidad Nacional Autónoma de México, Instituto de

846 Geología, Boletín 118, 81–111.

847

848 Rubatto D., 2002. Zircon trace element geochemistry; partitioning with garnet and the link

849 between U-Pb ages and metamorphism. Chem Geol 184, 123–38. 850

851 Singer, D.A., Berger, V.I., Moring, B.C., 2005. Porphyry copper deposits of the world:

852 database, map, and grade and tonnage models. U.S. Geological Survey Open-File Report,

853 2005-1060.

854

855 Silver, L.T., Chappell, B.W., 1988. The Peninsular Ranges Batholith—an insight into the

856 evolution of the Cordilleran batholiths of southwestern North America. Trans. R. Soc. Edinb.

857 Earth Sci. 79, 105–121.

858

859 Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a

860 two-stage model. Earth Planet. Sc. Lett. 26, 207–221.

861

862 Smoliar, M., Walker, R., Morgan, J., 1996. Re–Os of group IIA, IIIA, IVA and IVB iron

863 meteorites. Science. 271, 1099–1102.

864

865 Sonder, L.J., Jones, C.H., 1999. "Western United States extension: How the west was

866 widened.": Annu. Rev. Earth. Planet. Sci. 27, 417–462.

867

868 Teixeira-Correia, C., Kirk, J.D., Frick, L.R., 2007. The Re-Os isotopic system: geochemistry

869 and methodology at the Geochronological Research Center (CPGeo) of the University of São

870 Paulo, Brazil. Geologia USP. Série Científica. 7, 45–56.

871 872 Titley, S.R., 1982. Geologic setting of porphyry copper deposits, southeastern Arizona, in

873 Titley, S.R., (Ed.), Advances in the geology of the porphyry copper deposits in the

874 southwestern North America. Tucson Arizona, University of Arizona Press, p. 37–58.

875

876 Titley, S.R., 1993. Characteristics of porphyry copper occurrence in the American

877 Southwest, in Kirkham, R.V., Sinclair, W.D., Thorpe, R.I., Duke, J.M. (Eds.), Mineral deposit

878 modeling. Geol. Assoc. Canada Spec. Pap. 40, 433–464.

879

880 Valencia-Moreno, M., Iriondo, A., Gonzalez-Leon, C., 2006. Temporal constraints on the

881 eastward migration of the Late Cretaceous–early Tertiary magmatic arc of NW Mexico

882 based on new 40Ar/39Ar hornblende geochronology of granitic rocks. J. South Amer. Earth

883 Sci. 22, 22–38.

884

885 Valencia-Moreno, M., Ochoa-Landín, L., Noguez-Alcántara, B., Ruiz, J., Pérez-Segura, E.,

886 2007. Geological and metallogenetic characteristics of the porphyry copper deposits of

887 Mexico and their situation in the world context. Geol. Soc. Spec. Pap. 422, 433–458.

888

889 Valencia, V.A., Noguez-Alcántara, B., Barra, F., Ruiz, J., Gehrels, G., Quintanar, F., Valencia-

890 Moreno, M., 2006. Re-Os molybdenite and LA-ICPMS-MC U-Pb zircon geochronology for the

891 Milpillas porphyry copper deposit: insights for the timing of mineralization in the Cananea

892 District, Sonora, Mexico. Rev. Mex. Ciencias Geológicas, 23, 39–53.

893 894 Valentine, W.G., 1936. Geology of the Cananea Mountains, Sonora, Mexico. Geol. Soc. Am.

895 Bull. 47, 53–86.

896

897 Valenzuela-Navarro, L. C., Valencia-Moreno, M., Thierry-Calmus, Ochoa-Landín, L.,

898 González-León, C., 2005. Marco geológico del pórfido de molibdeno El Crestón, Sonora

899 central, México: Revista Mexicana de Ciencias Geológicas. 22, 345–357.

900

901 Varela, F.E., 1972. Tourmaline in the Cananea mining district, Sonora, Mexico. Unpub. M.S.

902 Thesis, University of California, Berkeley, 79 p.

903

904 Vega-Granillo, R., Calmus, T., 2003. Mazatan metamorphic core complex (Sonora, Mexico):

905 structures along the detachment fault and its exhumation evolution. J. South Amer. Earth

906 Sci. 16, 193–204.

907

908 Velasco, J.R., 1966. Geology of the Cananea district, in Titley, S.R., Hicks, C.L. (Eds.), Geology

909 of the Porphyry Copper Deposits, Southwestern North America. Tucson Arizona, University

910 of Arizona Press, pp. 245–249.

911

912 Vikre, P.G., Graybeal, F.T., Fleck, R.J., Barton, M.D., Seedorff, E., 2014. Succession of Laramide

913 Magmatic and Magmatic-Hydrothermal Events in the Patagonia Mountains, Santa Cruz

914 County, Arizona. Econ. Geol. 109, 1667–1704.

915 916 Virtue, T.L., 1996. Geology and supergene enrichment at the Cananea porphyry copper

917 deposit, Sonora, Mexico. M.S. Thesis, University of Texas, El Paso Texas, 197 p.

918

919 Wilkens, J., Jr., Heidrick, T. L., 1995. Post-Laramide extension and rotation of porphyry

920 copper deposits, southwestern United States, in Pierce, F.W. Bolm, J.G., (Eds.), Porphyry

921 copper deposits of the American Cordillera. Arizona Geological Society Digest, p. 109–127.

922

923 Wodzicki, W.A., 1995. The evolution of Laramide igneous rocks and porphyry copper

924 mineralization in the Cananea district, Sonora, Mexico. Unpublished Ph.D. dissertation,

925 University of Arizona, Tucson Arizona, 181 p.

926

927 Wodzicki, W.A., 2001. The Evolution of Magmatism and Mineralization in the Cananea

928 District, Sonora, Mexico, in Albinson, T. Nelson, C.E. (Eds.), New Mines and Discoveries in

929 Mexico and Central America. Econ. Geol. Spec. Pub. 8, 241–61.

930

931 932

933

934 Figure Click here to download Figure: Figure caption.docx

1 Figure caption

2 Fig. 1 Map showing the distribution of Laramide intrusive (black) and extrusive (grey)

3 rocks in Sonora. BC: Buenavista del Cobre mine; LC: La Caridad mine.

4

5 Fig. 2 Distribution of porphyry copper deposits along the Laramide belt of southwestern

6 North America and northwestern Mexico (modified after Valencia-Moreno et al., 2007).

7 Shaded area represents the Great Laramide porphyry copper cluster of Arizona, Sonora,

8 and New Mexico. Dashed line denotes the Cananea lineament (Hollister, 1978). Chrontours

9 according to Valencia-Moreno et al. (2006).

10

11 Fig. 3 Map showing the regional geological framework of the Cananea Mining district in

12 northeastern Sonora modified after Del Rio-Salas et al. (2013) and Servicio Geológico

13 Mexicano (Mexican Geological Survey). EP: El Pilar; LU: Lucy; TO: Toro; A: Alisos; MI:

14 Milpillas; LV: La Verde; MA: Mariquita; M: Maria; BC: Buenavista del Cobre; EA: El Alacrán.

15

16 Fig. 4 U–Pb weight average plots from the mineralizing porphyritic units from Buenavista

17 del Cobre porphyry copper deposit.

18

19 Fig. 5 U–Pb weight average plots from the mineralizing porphyritic unit from El Alacrán (a),

20 and El Pilar pluton (b-c).

21

22

23 24 Fig. 6. Diagram showing the southeast age progression of magmatic activity and

25 molybdenite mineralization. Grey bars indicate mineralizing pulses constrained by Re-Os

26 molybdenite data.

27

28 Fig. 7: Interpretative map of the main Tertiary extensional structures of the Cananea

29 mining district. The region is dissected by NNW-SSE normal faults, limiting horsts and

30 grabens, of which geographic distribution defines a NE-SW lineament between Imuris and

31 Milpillas (see text for explanation). The U-Pb and Re-Os isotopic ages shows two possible

32 progressions: 1) an ENE progression, following the classical eastward migration previously

33 documented for the whole Laramide magmatic arc (see text for more details); 2) a SE

34 progression along the Cananea lineament that may control locally the emplacement of the

35 ore bodies. The NNE-SSW dotted black lines are perpendicular to the ENE progression and

36 intercept the ore bodies dated by Re-Os method. The upper age (italic) corresponds to the

37 U-Pb zircon ages of the porphyritic intrusions, except for El Pilar Deposit. The lower age

38 (regular) corresponds to the molybdenite Re-Os ages. In the case of the Buenavista del

39 Cobre mine, the 60.8 Ma was chosen between three available ages.

40

41 Tables

42 Table 1. General geologic features of the Porphyry copper deposits from the Cananea district,

43 northwestern Mexico.

44

45 Table 2. Re-Os molybdenite data from the El Pilar, Mariquita, and Lucy copper deposits from

46 the Cananea district. 47

48 Table 3. U-Pb geochronologic analyses of granodiorite porphyry from Buenavista del Cobre

49 mine.

50

51 Table 4. U-Pb geochronologic analyses of granodiorite porphyry from Buenavista del Cobre

52 mine.

53

54 Table 5. U-Pb geochronologic analyses of quartz monzonite porphyry from Buenavista del

55 Cobre mine.

56

57 Table 6. U-Pb geochronologic analyses of monzodiorite porphyry from Buenavista del Cobre

58 mine.

59

60 Table 7. U-Pb geochronologic analyses of the mineralizing porphyry from El Alacrán PCD.

61

62 Table 8. U-Pb geochronologic analyses of the granodiorite from the El Pilar Cu deposit.

63

64 Table 9. U-Pb geochronologic analyses of the granodiorite from the El Pilar Cu deposit.

65

66 Figure 1 Click here to download Figure: Figure 1.pdf

Del Rio Salas et al., Fig. 1

32o Metamorphic core complex belt

USA MEX BC

LC

30o

Hermosillo Gulf of California

28o

Laramide rocks Plutonic Volcanic 100 km

114o 112o 110o Figure 2 Click here to download Figure: Figure 2.pdf

Del Rio-Salas et al., Fig. 2

114° Arizona 110° New Mexico

California

Tucson 32° 80

Sonora BC Gulf of LC California 70

Baja California

60 Chihuahua Hermosillo

28°

Baja California Sur

100 km Sinaloa Figure 3 Click here to download Figure:110°40'W Figure 3.pdf 110°30'W ´ N ' EP 3 5 km 0 1 1

1 °1 ° 0 1 'N 3

1 A 1 1TO MI

1 LU 11 0 °4 0 LV 'W 1 MA 1

M 1

3 4$OOXYLDOGHSRVLWV 1 °0 'N 74%DVLQILOOVHGLPHQWV 31 Cananea °0 75K\ROLWLFSRUSK\ULHV 'N 74XDUW]IHOGVSDUSRUSK\ULHV 1 7'DFLWLFSRUSK\ULHV BC E E E E 1 E E E E .77LQDMD&XLWDFDEDWKROLWK 1 E E E E .0HVD)RUPDWLRQ 11 0 °3 0 .0DULTXLWDGLDEDVH 'W .&DEXOORQD*URXS .9ROFDQRFODVWLFURFNV -7RUUHV\HQLWHVWRFN -6DQGVWRQHVDQGUK\ROLWLFIORZV 7RZQ 1 EA - " +HQULHWWD)RUPDWLRQ &RQWDFW

7U-(OHQLWD)RUPDWLRQ 3

0 °

1RUPDOIDXOW 5

3]8QGLIIHUHQWLDWHG3DOHR]RLF 3 0 '

0 N °5 )DXOW 0 Є%ROVD)RUPDWLRQ 'N D D D D D D D D 3F&DQDQHD*UDQLWH 1 0LQHSURVSHFW D D D D 110°20'W 110°10'W Figure 4 Click here to download high resolution image Figure 5 Click here to download high resolution image Figure 6 Click here to download high resolution image Figure 7 Click here to download Figure: Figure 7.pdf

Sierra San Antonio ARIZONA 111ºW Nogales 110º30’W 110ºW

SONORA Sierra Chivato Sierra Sierra San José EP

Sierra El Pinito ENE progression N

San Antonio Río San Pedro valley

MI

LU Milpillas - Milpillas Sierra Mariquita 0 10 km

MA Sierra ElenitaM

Sierra Los Ajos Cocóspera graben - Cocóspera Cuitaca - río Bacanuchi graben 31ºN Cananea Sierra Cuitaca Río Sonora graben

BC

SE progression Sierra Magdalena Sierra

Sierra El Manzanal Ímuris

EA Sierra Madera Sierra 63.8 Sierra Azul 61.8 63.8 60.8 63.0 59.2 74.6 73.9 62.7 57.8 60.4 60.8 Table 1

Table 1. Intrusive rocks Age Pre-min Porphyry (Ma) Ton Deposit name Metals Style Method Mineralogy (x106) Metal contents References

Buenavista py, cpy, mo, cc, 0.42% Cu, 0.008% Cu-Mo-Zn sw, b, sk gd, mz-di qz-feld 59.2-59.3 ± 0.3 Re-Os 7,140 1, 2, 3, 4 del Cobre co, en Mo, 0.58 gr/ton Ag, 0.012 gr/ton Au Milpillas Cu sw gd qz-feld 63.0-63.1 ± 0.4 Re-Os cpy, oxides 230 0.85% Cu 5, 6

Mariquita Cu-Mo sw, b gd, mz-di qz-feld 59.2-59.3 ± 0.3 Re-Os py, cpy, cc 100 0.48% Cu 7, 8, 9

María Cu-Mo sw, b gd qz-feld 60.4 ± 0.3 Re-Os py, cpy, mo 8.6 1.7% Cu, 0.1% Mo 1, 4

El Alacrán Cu-Mo sw, b gd qz-mz 60.8-60.9 ± 0.2 Re-Os py, cpy, cc, mo 2.4 0.35% Cu 1, 4, 7

Lucy Mo-Cu sw gd gd 61.6-61.8 ± 0.3 Re-Os mo, cpy - - 9

El Pilar Cu-Mo sw gd - 73.9 ± 0.4 Re-Os cpy, py, mo - - 9

Mineralization style: (sw) stockwork and veins; (sk) skarn; (b) breccia. Intrusive rocks: (qz-feld) quartz-feldespatic porphyry; (di) diorite; (mz) monzonite. Metallic mineralogy: (cc) chalcocite; (co) covellite; (cpy) chalcopyrite; (en) enargite; (mo) molybdenite; (py) pyrite. References: (1) Wodzicki, 2001; (2) Barton et al., 1995; (3) Singer et al., 2005; (4) Barra et al., 2005; (5) Valencia et al., 2006; (6) Noguez-Alcántara, 2008; (7) Pérez-Segura, 1985; (8) Del Rio Salas et al., 2013; (9) Present study. Table 2

Table 2.

Total Re Deposit Sample 187Re (ppm) 187Os (ppb) Age (Ma) (ppm) Mariquita Mari-1 83.7 52.6 51.6 59.3 ± 0.3 Mariquita Mari-2 373.5 234.8 231.6 59.2 ± 0.3 Lucy Lucy-1 51.55 32.41 33.28 61.6 ± 0.3 Lucy Lucy-2 47.2 47.2 29.7 61.8 ± 0.3 El Pilar Pilar-2 64.8 40.7 50.2 73.9 ± 0.4

Table 3

Table 3.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 384 2.2 2790 0.0091 6.5 58.2 3.8 2 865 2.2 2542 0.0094 2.1 60.4 1.2 3 1408 2.1 9186 0.0090 4.7 57.8 2.7 4 1812 2.2 11160 0.0091 2.3 58.1 1.3 5 659 1.5 2032 0.0094 5.6 60.5 3.4 6 795 1.7 3964 0.0091 1.4 58.7 0.8 7 702 2.1 7108 0.0095 2.5 60.8 1.5 8 968 2.4 15772 0.0094 2.8 60.4 1.7 9 1069 1.7 6016 0.0095 2.1 60.6 1.3 10 680 2.5 14452 0.0095 1.9 60.7 1.1 11 186 1.0 940 0.0090 1.9 57.5 1.1 12 877 1.7 5760 0.0095 0.9 61.0 0.5 13 911 2.2 4822 0.0097 1.1 61.9 0.7 14 883 2.0 4968 0.0095 0.7 61.1 0.4 15 905 2.3 4564 0.0095 4.0 60.7 2.4 16 358 1.8 1850 0.0096 2.4 61.3 1.5 17 935 2.1 6034 0.0096 1.0 61.5 0.6 18 584 2.6 2326 0.0095 2.8 61.2 1.7 19 324 2.1 3400 0.0096 1.6 61.6 1.0 20 766 2.3 4542 0.0094 1.3 60.2 0.8 21 1069 2.0 7408 0.0095 1.7 61.0 1.0 22 658 1.5 3612 0.0095 4.4 60.6 2.7

Table 4

Table 4.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 446 1.8 2680 0.0092 3.5 59.0 2.0 2 634 1.0 3684 0.0094 3.2 60.6 1.9 3 523 1.8 2482 0.0095 2.6 61.0 1.6 4 954 1.4 3958 0.0095 1.6 60.8 1.0 5 972 4.5 5264 0.0096 1.6 61.4 1.0 6 409 2.1 2496 0.0094 2.4 60.4 1.4 7 580 2.0 3636 0.0093 3.2 59.6 1.9 8 409 1.6 2218 0.0095 4.9 61.0 3.0 9 608 2.9 3656 0.0097 2.2 62.4 1.4 10 1060 0.9 4676 0.0092 3.8 59.2 2.2 11 589 2.1 2442 0.0094 2.9 60.4 1.7 12 1157 1.0 5008 0.0094 2.0 60.2 1.2 13 788 2.0 4618 0.0097 1.8 62.0 1.1 14 518 2.4 2418 0.0092 5.4 58.7 3.2 15 1739 0.9 7466 0.0095 2.0 60.8 1.2 16 551 1.6 1610 0.0090 1.8 57.6 1.0 17 545 2.0 3396 0.0095 2.0 61.2 1.2 18 424 1.9 2396 0.0094 2.7 60.2 1.6

Table 5

Table 5.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 673 1.9 4166 0.0095 1.7 61.1 1.0 2 665 2.0 2956 0.0091 2.2 58.4 1.2 3 452 1.5 2876 0.0096 1.8 61.6 1.1 4 433 1.9 1618 0.0095 5.2 61.0 3.1 5 273 2.4 2006 0.0097 2.8 62.4 1.7 6 357 2.1 1910 0.0094 2.1 60.0 1.2 7 702 1.3 3196 0.0096 1.2 61.3 0.7 8 449 1.8 2386 0.0094 2.2 60.4 1.3 9 308 2.2 1584 0.0094 2.9 60.0 1.8 10 407 2.1 2200 0.0098 2.1 62.8 1.3 11 1313 2.0 6102 0.0091 3.2 58.6 1.9 12 840 1.5 3222 0.0095 2.9 60.7 1.7 13 381 2.1 2052 0.0095 3.7 60.9 2.2 14 369 2.2 2500 0.0098 2.4 62.8 1.5 15 482 1.9 3824 0.0096 0.6 61.4 0.4 16 741 1.7 3552 0.0093 2.1 59.9 1.3

Table 6

Table 6.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 1774 1.7 7764 0.0094 1.6 60.4 1.0 2 118 1.0 418 0.0090 7.1 57.9 4.1 3 949 1.5 3976 0.0094 1.6 60.5 1.0 4 509 1.7 2760 0.0094 2.2 60.3 1.3 5 1918 2.1 7020 0.0089 2.3 57.4 1.3 6 1448 2.3 2590 0.0090 2.6 58.0 1.5 7 1814 2.5 7160 0.0090 2.6 57.9 1.5 8 1842 2.2 9410 0.0092 2.7 58.8 1.6 9 600 1.7 1978 0.0091 2.0 58.3 1.2 10 2018 2.0 7184 0.0090 2.3 58.1 1.3 11 2347 1.9 8246 0.0087 1.9 55.7 1.0 12 2156 1.8 7340 0.0091 1.1 58.3 0.6 13 1397 2.6 5734 0.0092 1.8 59.1 1.1 14 1379 2.2 3496 0.0094 2.0 60.5 1.2 15 1443 2.1 7160 0.0094 2.0 60.1 1.2 16 1698 2.2 8870 0.0094 1.7 60.3 1.0 17 1525 2.8 6376 0.0093 2.0 59.5 1.2 18 2106 2.0 16426 0.0092 1.8 58.9 1.0 19 1701 1.7 4758 0.0088 1.6 56.3 0.9 20 1265 3.1 11920 0.0090 1.7 58.1 1.0 21 1280 2.6 5120 0.0092 3.4 59.0 2.0 22 1458 2.8 7142 0.0094 2.9 60.2 1.7 23 1031 2.0 5168 0.0090 3.5 57.6 2.0 24 263 1.6 1170 0.0092 5.5 59.3 3.3

Table 7

Table 7.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 889 1.9 3914 0.0091 1.3 58.5 0.7 2 340 1.7 1894 0.0090 2.6 57.4 1.5 3 899 1.8 3478 0.0091 1.8 58.3 1.1 4 442 2.0 1792 0.0090 4.7 57.7 2.7 5 263 3.1 1298 0.0090 4.8 57.8 2.8 6 611 2.2 1482 0.0090 1.7 57.9 1.0 7 1621 1.9 7274 0.0089 3.0 57.2 1.7 8 992 2.7 13054 0.0089 1.6 56.8 0.9 9 1419 2.1 2182 0.0089 2.8 56.8 1.6 10 1236 2.3 3202 0.0090 1.0 57.7 0.6 11 866 2.8 4352 0.0090 0.8 58.0 0.5 12 1478 2.3 5096 0.0087 2.4 56.1 1.3 13 653 1.9 1242 0.0089 2.0 56.8 1.1 14 1476 2.0 6416 0.0091 1.9 58.4 1.1

Table 8

Table 8.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 274 4.1 2018 0.0116 2.7 74.2 2.0 2 140 4.1 1450 0.0115 4.1 73.6 3.0 3 382 3.7 5250 0.0116 1.8 74.5 1.4 4 281 4.0 1524 0.0113 5.2 72.2 3.8 5 535 3.0 17918 0.0113 1.5 72.6 1.0 6 416 3.5 3016 0.0114 3.2 73.0 2.3 7 191 4.3 10632 0.0113 2.9 72.6 2.1 8 296 3.9 2302 0.0114 5.0 73.1 3.6 9 306 2.5 2206 0.0117 1.3 75.1 1.0 10 327 1.9 2918 0.0113 2.3 72.7 1.7 11 426 2.7 3150 0.0118 1.8 75.5 1.3 12 304 4.2 1994 0.0120 1.8 76.6 1.3 13 468 2.9 3416 0.0115 3.0 73.9 2.2 14 495 2.1 3052 0.0115 1.2 73.9 0.9 15 179 3.9 5262 0.0118 2.7 75.4 2.0 16 220 2.6 2060 0.0118 1.9 75.7 1.4 17 309 3.0 2352 0.0118 3.3 75.8 2.5 18 205 3.5 956 0.0114 2.5 72.8 1.8 19 200 3.4 2148 0.0116 4.6 74.2 3.4 20 233 3.6 1864 0.0120 4.2 77.1 3.2 21 451 2.5 9820 0.0114 2.0 73.2 1.5 22 399 3.2 1664 0.0119 2.0 76.0 1.5 23 293 2.9 2434 0.0118 2.7 75.6 2.1 24 282 2.1 1592 0.0120 1.6 76.9 1.2 25 273 3.1 1572 0.0116 2.2 74.1 1.6 26 656 2.6 4310 0.0117 4.6 75.2 3.4 27 310 3.3 2194 0.0117 1.9 74.8 1.4

Table 9

Table 9.

Analysis U (ppm) U/Th 206Pb/204Pb 206Pb*/238U ± (%) 206Pb*/238U* ± (Ma) ratio age 1 346 2.6 2574 0.0119 3.6 76.2 2.7 2 480 3.2 4420 0.0118 2.9 75.4 2.2 3 386 2.4 3408 0.0114 2.0 72.8 1.4 4 350 2.9 2112 0.0118 4.1 75.9 3.1 5 672 2.9 4248 0.0119 1.5 76.4 1.2 6 300 1.6 2474 0.0119 2.4 76.3 1.8 7 606 3.6 4406 0.0114 2.6 73.0 1.9 8 349 2.4 7646 0.0118 1.4 75.7 1.0 9 366 3.2 2664 0.0114 3.1 72.8 2.3 10 898 2.7 6128 0.0116 2.5 74.3 1.9 11 490 2.7 3156 0.0115 1.4 73.5 1.0 12 517 3.0 2650 0.0119 4.8 76.3 3.6 13 576 2.8 5672 0.0117 5.1 75.3 3.8 14 667 1.6 4384 0.0115 1.7 74.0 1.3 15 548 1.9 4542 0.0114 3.4 73.3 2.5 16 535 2.6 5068 0.0122 0.9 78.2 0.7 17 567 3.4 3996 0.0114 3.7 73.2 2.7 18 890 3.8 1974 0.0113 2.7 72.7 2.0 19 702 3.1 4426 0.0115 1.1 73.6 0.8 20 724 2.8 5052 0.0113 3.2 72.6 2.3 21 584 2.8 4570 0.0121 3.0 77.4 2.3 22 703 2.9 4682 0.0117 2.1 75.0 1.5 23 405 2.6 3154 0.0116 2.8 74.5 2.0 24 839 2.7 6498 0.0119 1.5 76.3 1.1 25 626 1.8 6278 0.0115 3.4 73.5 2.5 26 696 2.4 6156 0.0118 3.9 75.7 3.0 27 505 2.5 3124 0.0119 2.6 76.5 2.0 28 522 3.1 6762 0.0116 2.0 74.4 1.4 29 577 4.2 2490 0.0119 1.3 76.3 1.0 30 400 2.2 2374 0.0115 3.3 73.6 2.4 31 439 3.9 3534 0.0115 2.5 74.0 1.8