Holford et al. Post-Caledonian exhumation of northwest

1 Multiple post-Caledonian exhumation episodes across northwest Scotland revealed by

2 apatite fission track analysis

3

4 Simon P. Holford1, Paul F. Green2, Richard R. Hillis1, John R. Underhill3, Martyn S. Stoker4

5 & Ian R. Duddy2

6

7 1Australian School of Petroleum, Centre for Tectonics, Resources and Exploration (TRaX), University of Adelaide, SA 5005, Australia

8 ([email protected])

9 2Geotrack International Pty Ltd, 37 Melville Road, Brunswick West, Victoria 3055, Australia

10 3School of Geosciences, The University of Edinburgh, Grant Institute of Earth Science, The King’s Buildings, West Mains Road, Edinburgh

11 EH9 3JW,

12 4British Geological Survey, Murchison House, West Mains Road, Edinburgh EH9 3LA, United Kingdom

13

14 Abstract

15 The post Caledonian exhumation history of northwest Scotland is a controversial issue, with

16 some studies advocating largely continual emergence while others suggest dominantly early

17 Palaeogene plume-driven exhumation. AFTA data in samples of Precambrian basement and

18 Permian-Cretaceous sediments from onshore and onshore reveal multiple phases of post-

19 Caledonian cooling, viz: Triassic (beginning 245-225 Ma), Cretaceous (140-130 Ma; 110-90

20 Ma) and Cenozoic (65-60 Ma; 40-25 Ma; 15-10 Ma), all of which are interpreted at least in

21 part as recording exhumation. Basement and sedimentary cover rocks display similar thermal

22 histories, emphasising the regional nature of these episodes and implying that sedimentary

23 outliers represent the remnants of previously more extensive sequences. Significant

24 thicknesses of Jurassic rocks may once have covered northwest Scotland. Palaeocene

25 palaeothermal effects are most pronounced in the vicinity of igneous centres, probably

26 reflecting combined effects of heating by elevated heat flow, deeper burial and hydrothermal

27 activity. Most of the region underwent km-scale Neogene exhumation. Contrary to the

1 Holford et al. Post-Caledonian exhumation of northwest Scotland

28 common assumption of monotonic cooling and denudation histories, integration of geological

29 evidence with AFTA data defines an episodic thermal history involving repeated cycles of

30 burial and exhumation. We suggest that onshore passive margins and continental interiors

31 may also best be characterised by similar histories.

32

33

34

35

36 Abstract

37 The origin and development of the high topography bordering the Atlantic margin of NW

38 Europe has long been the subject of debate (Geike, 1901; George, 1966; Doré et al., 2002).

39 The mountains of Scotland and western Scandinavia are characterized by rugged topography

40 and elevations that exceed 1.3 km and 2.4 km respectively and largely comprise rocks that

41 were formed and/or deformed during the Caledonian Orogeny. Numerous lines of evidence

42 indicate that the present-day topography of both regions was initiated by tectonic uplift that

43 began during the Cenozoic or late Mesozoic and which also affected peripheral sedimentary

44 basins (Doré et al., 2002). These include; the transition from deposition of shallow-marine

45 carbonates in the late Cretaceous to rejuvenated clastic sedimentation in the early Palaeogene

46 (Doré et al., 2002)); the easterly tilt of Britain (Brodie & White, 1994); identification of

47 Palaeocene, Eocene and Pliocene-Pleistocene prograding and locally heavily incised shelf-

48 slope wedges of clastic sediments offshore Scotland and Norway (Stoker et al., 1993, 2010;

49 Underhill, 2001); apatite fission track, vitrinite reflectance and sonic velocity data which

50 record deeper burial of Cenozoic-Mesozoic sediments and Palaeozoic and older basement

51 rocks along the Atlantic margin prior to their uplift and erosion (Japsen et al., 2007; Hillis et

52 al., 2008); and, geomorphological evidence for elevation of low-relief erosion surfaces and

2 Holford et al. Post-Caledonian exhumation of northwest Scotland

53 landscape elements formed at or near sea level during the Mesozoic and Palaeogene and

54 subsequently uplifted (Japsen et al., 2002).

55

56 Despite this evidence for tectonically driven uplift of the Atlantic margin during the late

57 Mesozoic and Cenozoic, it has been recently proposed (Nielsen et al., 2009) that the present-

58 day topography of western Scandinavia and the represents remnant

59 topography from the Caledonian Orogeny that survived post-orogenic extensional collapse

60 and erosion. Nielsen et al. (2009) suggest that many of the observations listed above that have

61 been used to argue for Cenozoic tectonic uplift can instead be interpreted in terms of

62 protracted exhumation maintained by climatically-controlled erosion rates. This hypothesis

63 implies that the Scottish Highlands and western Scandinavia have been positive topographic

64 features for the past ~400 million years, experiencing no significant burial and receiving little

65 or no Upper Palaeozoic-Cenozoic sediments (cf. Macdonald et al., 2007).

66

67 Recent apatite fission track and (U-Th)/He data from basement samples in the Scottish

68 Highlands have been interpreted in terms of thermal histories dominated by continuous,

69 monotonic cooling and exhumation over the past ~400 Myr (Persano et al., 2007) and thus

70 support the idea of continuous, post-Caledonian emergence (cf. Watson, 1985). Such

71 monotonic cooling of basement rocks is inconsistent with the local geological record

72 however, because outliers of Devonian, Carboniferous, Permian-Triassic, Jurassic and

73 Cretaceous sediments overlying Neoproterozoic basement (e.g. Moine Supergroup,

74 Torridonian) are found along the coast of northwest Scotland (Fig. 1). In addition, thick

75 sequences of mainly Triassic-Jurassic age sediments are preserved in offshore basins of the

76 Malin-Hebrides sea area (Fyfe et al., 1993), whilst the overstep of Old Red Sandstone onto

3 Holford et al. Post-Caledonian exhumation of northwest Scotland

77 late Caledonian granites indicates that Caledonian topography had largely been worn down

78 by rapid uplift and erosion during the early Devonian (Hall, 1991).

79

80 Conversely, other workers have claimed that Scotland has experienced post-Caledonian

81 tectonically-driven uplift that was dominantly early Palaeogene in age and driven by

82 processes associated with the Iceland mantle plume, such as igneous underplating or

83 convectively-supported dynamic uplift (White & Lovell, 1997; Jones et al., 2002; Shaw

84 Champion et al., 2008). However, recent fission-track studies from the southern British Isles

85 have revealed a more complex Mesozoic-Cenozoic uplift history, with major phases of km-

86 scale uplift and exhumation occurring during the early Cretaceous and Neogene in addition to

87 the early Palaeogene (Holford et al., 2005; Hillis et al., 2008; Holford et al., 2009a, b, c).

88

89 The aim of this study is to test these different models for the post-Caledonian tectonic

90 evolution of northwest Scotland using a new apatite fission-track analysis (AFTA) database

91 comprising onshore and offshore samples from basement (i.e. Lower Palaeozoic and older)

92 and from the Permian-Cretaceous sediments deposited in overlying sedimentary basins. The

93 offshore database includes samples collected over a range of depths from exploration wells

94 located in the Sea of the Hebrides (134/5-1) and West (202/19-1) basins that provide

95 direct constraints on palaeogeothermal gradients and allow estimates of amounts of missing

96 section removed during exhumation. Thermal history solutions extracted from our data show

97 consistency between samples from basement and cover sequences and show that the

98 Mesozoic sediments deposited around northwest Scotland have been heated to temperatures

99 exceeding 100°C during burial by additional sedimentary sequences. The preserved

100 sediments are thus interpreted as remnants of deposits that were originally far more extensive,

101 raising the possibility that the Highlands were buried beneath a Mesozoic sedimentary cover.

4 Holford et al. Post-Caledonian exhumation of northwest Scotland

102 The additional sedimentary sequences were removed during a series of Cretaceous-Cenozoic

103 exhumation episodes, whose timing is constrained by both AFTA data and unconformities

104 that separate the sedimentary units. Our results indicate that interpretations involving

105 continual post-Caledonian emergence cannot be sustained, and that the post-400 Ma history

106 of northwest Scotland has been characterised by repeated cycles of sedimentation and

107 exhumation, suggesting that post-Caledonian events are largely responsible for the

108 physiography of northwest Scotland.

109

110 Post-Caledonian geological history of northwest Scotland

111 The Caledonian orogeny in northwest Scotland culminated during the Lower Devonian (Fyfe

112 et al., 1993). The onset of extensional tectonics in the Middle Devonian facilitated the

113 collapse of the Caledonian mountain belt (Stoker et al., 1993). Thick sequences of Old Red

114 Sandstone (ORS) continental sediments and lavas occur within the Midland Valley, southeast

115 of the Boundary Fault (Fyfe et al., 1993) and large areas of Devonian sedimentary

116 rocks in northeast Scotland are remnants of the originally more extensive Orcadian Basin

117 (Hillier and Marshall, 1992) (Fig. 1). Limited ORS siliciclastic and volcanic outcrops of

118 Lower Devonian age around Oban and in the Firth of Lorne are interpreted to be the

119 preserved remnants of a formerly more extensive ORS cover (Fyfe et al., 1993).

120 Carboniferous rocks are of limited exposure in the Scottish Highlands with the exception of

121 ~100 m of Carboniferous sandstone preserved on (Fyfe et al., 1993), although

122 considerably thicker sequences (up to 3 km) of marine and nonmarine Carboniferous

123 sediments are preserved within the Midland Valley Basin (Underhill et al., 2008).

124

125 Permian-Triassic continental redbeds are widely distributed in the basins offshore northwest

126 Scotland (Fig. 1), reaching thicknesses of ~3 km in the North Minch Basin (Fyfe et al., 1993).

5 Holford et al. Post-Caledonian exhumation of northwest Scotland

127 Upper Triassic sediments are exposed on Skye, , Mull, Morvern, Applecross,

128 Gruinard Bay and on northeast Lewis where ~1.2 km of Permian-Triassic sandstones and

129 conglomerates rest with unconformable or faulted contact against Lewisian gneisses (Fyfe et

130 al., 1993). These basins also contain thick sequences of marine and nonmarine Jurassic

131 deposits, and there are onshore outcrops along the northwestern coast of Scotland and on

132 Skye and where the Jurassic succession reaches 450 m (Fyfe et al., 1993). Based on

133 detailed subsidence analyses of preserved Triassic-Jurassic rocks in the ,

134 Morton (1987) proposed that the initial phase of crustal stretching responsible for the

135 Hebridean basin system commenced during the late Triassic (~210 Ma).

136

137 Cretaceous rocks are less widely distributed (Fig. 1), but this is considered to be due

138 Cenozoic uplift and erosion rather than non-deposition (Hallam, 1983). Upper Cretaceous

139 sediments are preserved on Morvern, including glauconitic and pure silica sandstones

140 overlain by silicified Chalk (Fyfe et al., 1993). The Cretaceous rocks on Morvern rest

141 unconformably on Triassic-Jurassic strata, implying an intervening period of erosion,

142 probably during the early Cretaceous (Hallam, 1983). There are minor outcrops of Upper

143 Cretaceous strata on Skye, Mull, Eigg and Arran but no rocks of this age have been proven

144 immediately offshore (Fyfe et al., 1993).

145

146 Widespread volcanicity accompanied continental breakup between Greenland and northwest

147 Europe in the late Palaeocene-early Eocene along the North Atlantic margins (Emeleus and

148 Bell, 2005). Most volcanicity in the British Isles was concentrated in northwest Scotland,

149 with major igneous centres at Mull, Ardnamurchan, Rum and Skye (Emeleus and Bell, 2005).

150 This igneous activity has been attributed to the proto-Iceland mantle plume, and petrological

151 data suggest that a significant proportion (up to 70%) of the melt generated around the British

6 Holford et al. Post-Caledonian exhumation of northwest Scotland

152 Isles was trapped at depth, leading to underplating of the lower crust, which in turn would

153 have resulted in surface uplift (Jones et al., 2002). However, the depth to the seismic

154 reflection Moho beneath northwest Scotland is low compared to southeast Scotland (varying

155 from 25 to 29 km beneath basins and basement) (Chadwick and Pharaoh, 1998) and is

156 inconsistent with the crust having been significantly thickened by igneous material.

157

158 Climatic deterioration during the Quaternary profoundly affected the landscape of the

159 Scottish Highlands, with decreasing temperatures permitting the growth of glaciers on

160 uplifted terrain which in turn promoted widespread erosion and deposition of glacial

161 glaciomarine sediments offshore northwest Scotland (Stoker et al., 1993).

162

163 AFTA methodology and database

164 The methodological and analytical aspects of AFTA are well established (e.g. Green et al.,

165 2002). AFTA utilizes the radiation damage trails (fission tracks) created by spontaneous

166 fission of 238U within the crystal lattice of apatite, which is a common detrital mineral in most

167 sandstone and occurs as an accessory phase in plutonic and high grade metamorphic rocks.

168 By counting the number tracks in a polished and etched grain surface and measuring the

169 uranium content, a fission track age can be calculated, which in the absence of other factors

170 would provide a direct measure of the time over which tracks have accumulated. Fission

171 track lengths form within a narrow range (~16 μm), but are progressively shortened as

172 radiation damage is repaired at a rate that increases with temperature (annealing). Shortening

173 leads to a reduction in the number of tracks that can intersect a polished surface, and above

174 110°C all damage is repaired and no tracks are preserved. A measured fission track age thus

175 represents a balance between production of tracks and loss by annealing, and must be

176 assessed in tandem with track length data.

7 Holford et al. Post-Caledonian exhumation of northwest Scotland

177 We present AFTA data from 78 samples collected from onshore and offshore northwest

178 Scotland (Fig. 1, Table 1). Our dataset includes 42 samples from onshore outcrops, 21

179 samples from British Geological Survey (BGS) offshore shallow boreholes and drill cores,

180 and 15 samples from two wells (202/19-1 in the West Orkney Basin (WOB) and 134/5-1 in

181 the Sea of Hebrides Basin (SOHB)). The samples cover a broad range of stratigraphic levels,

182 from Proterozoic (Torridonian and Moine) and Archean (Lewisian) basement to Phanerozoic

183 sedimentary cover rocks of Permian, Triassic, Jurassic and Cretaceous age. Data quality is

184 generally excellent, with the majority of samples yielding 20 or more single grain ages and

185 many samples yielding 100 track lengths (Table 1).

186

187 The majority of measured fission track ages are younger than the respective stratigraphic ages

188 (Table 1). Overall there is a consistent trend between fission track age and mean track length,

189 with most fission track ages in the range of 150-300 Ma regardless of the depositional or

190 metamorphic age of the host rock, and mean track lengths (MTLs) in the range 11-13 µm

191 (Fig. 2). Samples from Skye yield much younger fission track ages between 50 and 70 Ma,

192 with mean track lengths around ~13-14 µm. The main trend in Fig. 2 is very different from

193 the simple “boomerang trend” shown by samples that have undergone a single heating event

194 in which different samples have undergone varying degrees of resetting (Green, 1986), and

195 suggests that the region has undergone a relatively complex history involving a series of

196 resetting events.

197

198 With the exception of one sample (GC503-151), the fission track ages of all samples from

199 northwest Scotland are younger than 400 Ma and the overwhelming majority are younger

200 than 300 Ma i.e. at least 100 Myr younger than the latest stages of Caledonian Orogeny. This

201 immediately indicates that almost all basement and early Palaeozoic samples reached

8 Holford et al. Post-Caledonian exhumation of northwest Scotland

202 palaeotemperatures sufficient to produce severe age reduction (i.e. ~100°C or above) during

203 post-Caledonian events. Moreover, Proterozoic and Achaean basement samples tend to yield

204 similar fission track ages and length distributions to nearby Phanerozoic cover samples

205 (Table 1 and Fig. 2), suggesting that basement and cover sequences in northwest Scotland

206 have undergone similar thermal histories.

207

208 Thermal history interpretation of AFTA data

209 Basic principles

210 Extraction of thermal history information from AFTA data begins with construction of a

211 Default Thermal History (DTH), derived from the burial history defined by the preserved

212 sedimentary section and the present-day thermal gradient. For a sedimentary rock, this is the

213 history which would apply if the sample has never been any hotter than the present-day

214 temperature at any time since deposition. For basement samples, a similar approach can be

215 adopted using the age of the oldest overlying sedimentary units, which in this study we have

216 assumed to be Devonian if no sedimentary units are present. If the DTH can explain the

217 AFTA data, then it is not possible to extract further thermal history information. If the AFTA

218 data show a greater degree of annealing (i.e. fission track age and/or track length reduction)

219 than expected from the DTH, then the sample must have been hotter in the past and

220 information on the magnitude and timing of heating and cooling events can be extracted from

221 the data. All samples analysed in this study show a greater degree of annealing (i.e. fission

222 track age and/or length reduction) than that expected from the DTH , and have therefore been

223 much hotter in the past.

224

225 To extract thermal history information from AFTA data, we use a kinetic model that makes

226 full quantitative allowance for the effects of Cl content on fission-track annealing rates

9 Holford et al. Post-Caledonian exhumation of northwest Scotland

227 (Green et al., 2002). Because maximum palaeotemperatures are the key factor that dominates

228 AFTA data, this technique reveals little information on the thermal history prior to the onset

229 of cooling. Therefore, we do not attempt to constrain the entire thermal history of each

230 sample but focus on the key aspects of the thermal history that control the fission track age

231 and length distribution i.e. the maximum palaeotemperature of each sample, and the time at

232 which cooling from that palaeotemperature began.

233

234 By modelling expected AFTA parameters resulting from a range of possible thermal

235 histories, we use maximum likelihood theory similar to that described by Gallagher (1995) to

236 define the range of values of maximum palaeotemperature and the onset of cooling giving

237 predictions which match the measured data within 95% confidence limits. Palaeotemperature

238 estimates derived using this approach usually have an absolute uncertainty of better than

239 ±10°C. Many of the samples analysed in this study require at least two episodes of heating

240 and cooling to explain all facets of the data. AFTA can provide constraints on two episodes

241 provided the magnitude and timing of the palaeothermal maximum and the subsequent

242 (lower) peak palaeotemperature are sufficiently separated. In some cases (e.g. Turner et al.,

243 2008) it is possible to resolve three discrete episodes in data from a single sample. This is

244 most likely when the first episode involves a maximum palaeotemperature sufficient to

245 totally anneal all tracks (typically >110°C), followed by a subsequent peak around 90- 100°C

246 which leads to shortening of tracks formed after the initial cooling to a mean length of ~10

247 μm. Finally, cooling to a low temperature is followed by reheating to ~70°C, sufficient to

248 reduce track lengths formed after the second episode to ~12-13 μm.

249

250 Table 2 summarises estimates of peak palaeotemperatures reached by each sample and the

251 time at which cooling from these palaeotemperatures began.

10 Holford et al. Post-Caledonian exhumation of northwest Scotland

252 West Orkney Basin

253 Of six Permian-Lower Jurassic samples from offshore shallow boreholes in the WOB, most

254 require two discrete episodes of heating and cooling (Table 2); an early episode involving

255 cooling from palaeotemperatures typically >90°C beginning at some time during the mid-

256 Jurassic to late Cretaceous (180 to 85 Ma based on combined results from all samples; Fig.

257 3). These palaeotemperatures imply significant depths of burial for any reasonable

258 palaeogeothermal gradient. This was followed by reheating and a second cooling episode

259 (from palaeotemperatures typically between 50- 70°C; Table 2) that commenced during the

260 mid-late Cenozoic (between 30 and 15 Ma; Fig. 3). These results are interpreted in terms of

261 burial of the preserved WOB succession by late Triassic-Jurassic sediments that were

262 removed by exhumation that occurred in two stages (mid-Jurassic to late Cretaceous, and

263 mid-late Cenozoic)

264

265 Exploration well 202/19-1

266 WOB well 202/19-1 was drilled in the hangingwall of the Solan Bank High Fault (Evans,

267 1997) and encountered ~3 km of Upper Permian-Lower Triassic sediments beneath a thin

268 veneer of Quaternary deposits (Stoker et al, 1993). AFTA results indicate that all nine

269 samples reached temperatures of ~100°C or above prior to cooling that began during the late

270 Jurassic to late Cretaceous (Table 2), similar to results from WOB boreholes. The consistent

271 timing of cooling based on overlap for all samples is 110 to 90 Ma (Fig. 3). A later episode of

272 cooling beginning during the late Cretaceous-Cenozoic is also required for most samples,

273 with the consistent overlap suggesting an onset beginning during the mid-late Cenozoic (55

274 to 10 Ma).

275

11 Holford et al. Post-Caledonian exhumation of northwest Scotland

276 The variation of palaeotemperature with depth for the two separate episodes (Fig. 4b)

277 suggests that both can be satisfied by linear palaeotemperature profiles, with gradients close

278 to the present-day gradient of ~25.3°C km-1. This indicates that these palaeotemperatures

279 record heating caused by deeper burial and subsequent cooling caused by exhumation (cf.

280 Green et al., 2002). Assuming a palaeosurface temperature of 5°C and palaeogeothermal

281 gradients equivalent to the present-day gradient, early Cretaceous palaeotemperatures are

282 consistent with deeper burial by 3.3-3.9 km of additional Triassic-Lower Cretaceous section,

283 whilst the mid-late Cenozoic palaeotemperatures are consistent with an additional 1.2-1.7 km

284 of section (Fig. 4c). Higher surface temperatures would require correspondingly lower

285 amounts of additional section, while higher palaeogeothermal gradients (within the ranges

286 allowed by the constraints on each episode) would require lower amounts of additional

287 section.

288

289 Evans (1997) estimated the amount of deeper burial of the Permian-Triassic section at

290 202/19-1 by comparing Triassic shale sonic velocities against published velocity-depth

291 relationships for normally compacted shales. By comparison with the burial curves of Marie

292 (1975) and Magara (1976), Evans (1997) estimated gross exhumation of the Permian-Triassic

293 section at 1.85 km. These velocity-depth relationships overestimate the increase of velocity

294 with depth in comparison to more recent trends for Triassic shales from the British Isles (e.g.

295 Japsen 2000). The value reported by Evans (1997) may thus underestimate the amount of

296 gross exhumation by several hundred metres. The amount of exhumation indicated by sonic

297 data is lower than that indicated by early Cretaceous AFTA palaeotemperatures assuming a

298 geothermal gradient of 25.3°C km-1 (i.e. 3.3-3.9 km) but is reasonably close to the maximum

299 likelihood value of additional section (2.31 km) (Fig. 4c). The slightly lower values of

12 Holford et al. Post-Caledonian exhumation of northwest Scotland

300 additional section indicated by the sonic velocity data could also indicate elevated basal heat

301 flow during the early Cretaceous.

302

303 Based on AFTA and sonic data from 202/19-1 and AFTA data from shallow boreholes we

304 suggest that up to 2-3 km of Upper Triassic-Lower Cretaceous sediments were deposited in

305 parts of the WOB prior to early Cretaceous (110-90 Ma) exhumation. Although some

306 samples are from footwall blocks, many others (including those from 202/19-1) are from

307 hangingwalls which implies a regional cause for exhumation as opposed to localized

308 exhumation of footwalls during Cretaceous normal faulting and extension.

309

310 Hebridean region

311 We analysed 22 samples from the North Minch Basin (NMB), and outcrops

312 along the northwest coast of Scotland. Most require two palaeothermal episodes, and thermal

313 history solutions for Permian and younger samples are generally similar to those resolved for

314 basement samples (Table 2). When combined, results from this region reveal up to four

315 separate cooling episodes during the Mesozoic-Cenozoic (Fig. 3).

316

317 Several of the Archean and Proterozoic samples, notably those in the footwalls of the NMB

318 from Lewis and Harris and the northwest coast of Scotland, require Permian-Triassic cooling

319 from temperatures generally >100°C (Table 2). An onset of cooling beginning between 225

320 and 190 Ma satisfies all samples that show evidence for this episode (Fig. 3).

321

322 Most samples from both basement and cover rocks in the Hebridean region show evidence

323 for cooling from palaeotemperatures of up to ~105°C during the Jurassic to mid-Cretaceous

324 (Table 2). Precise constraints on the timing of this episode are often broad in individual

13 Holford et al. Post-Caledonian exhumation of northwest Scotland

325 samples suggesting that Jurassic-mid Cretaceous cooling could result from several unresolved

326 events, but Permian-Lower Cretaceous sedimentary samples from shallow boreholes in the

327 Hebrides Shelf (GC523-47 to 52) define consistent evidence for cooling from ~90-95°C

328 beginning during the early Cretaceous (140 to 130 Ma) (Fig. 3).

329

330 Likewise, most samples from the Hebridean region reveal evidence for elevated

331 palaeotemperatures during the late Cretaceous-Cenozoic. Timing constraints are quite broad

332 for some samples, which is again suggestive of several distinct phases of cooling that are

333 closely spaced in temperature and time. It is possible to identify discrete late Cretaceous-mid

334 Cenozoic and late Cenozoic cooling episodes in data from Permian-Triassic sample GC523-

335 11 from NMB borehole 71/11. AFTA suggests that this sample cooled from 95-100°C

336 between 75 and 30 Ma, with further cooling from 55-80°C between 15 Ma and the present-

337 day. In Table 2 we have assigned late Cretaceous-Cenozoic cooling observed in individual

338 episodes to discrete late Cretaceous-mid Cenozoic (75 to 40 Ma) and late Cenozoic (15 to 0

339 Ma) episodes, with the former more prevalent in basement samples on the flanks of the NMB

340 (e.g. GC503-184) and the latter clearly resolved in samples from the Hebrides Shelf (e.g.

341 GC523-49, 56). However, we acknowledge that other interpretations of these results are

342 possible given the broad timing constraints for some samples.

343

344 Exploration well 134/5-1

345 We analysed eight AFTA samples from exploration well 134/5-1, located in the SOHB ~50

346 km southwest of Skye (Fig. 1). The well intersected ~2340 m of Upper Triassic-Lower

347 Jurassic section intruded by numerous Palaeocene minor sills and dykes (i.e. <2 m thick). We

348 also analysed 29 vitrinite reflectance (VR) samples (predominantly from the Jurassic section),

14 Holford et al. Post-Caledonian exhumation of northwest Scotland

349 using the kinetic model of Burnham and Sweeney (1989) to convert observed Romax values

350 to palaeotemperatures (using an assumed heating rate of 1°C Ma-1).

351

352 BHT data indicate a present-day geothermal gradient of 47.5°C km-1, but the DTH calculated

353 on this basis for sample GC369-6 predicts total annealing, whereas this sample contains

354 tracks and gives a finite (though very young) fission track age (Fig. 5a). In samples GC369-3

355 and 4 a similar comment applies to apatites containing <0.2 wt% Cl. In addition, GC369-4

356 and 5 show less length reduction than expected from the DTH. These aspects of the AFTA

357 data suggest that present-day temperatures based on a gradient of 47.5°C km-1 are too high,

358 and are more consistent with a gradient of ≤43°C km-1. We have used this upper limit as the

359 basis for construction of the DTH used to extract thermal histories from AFTA, and on this

360 basis six samples show clear evidence of having been hotter in the past (Table 2).

361

362 Combining timing constraints derived from AFTA samples with maximum

363 palaeotemperatures derived from VR data (Fig. 5b), indicates that two distinct Cenozoic

364 cooling episodes have affected the preserved section (Table 2). VR-derived

365 palaeotemperatures show fluctuations over narrow depth intervals around intrusions

366 superimposed on a “background” linear profile (Fig. 5b). The rapid fluctuations clearly

367 reflect contact heating around the igneous intrusions. AFTA data in GC369-2 is also

368 interpreted to reflect contact heating, and a similar conclusion applies to the

369 palaeotemperature constraint from GC369-1, as it plots above the main background trend

370 defined by VR and other AFTA samples (Fig. 5b). AFTA data from the Triassic section

371 (>1.75 km depth) define mid-Cenozoic cooling that began between 45 and 20 Ma (Table 2).

372 Lower limits to the maximum palaeotemperature in this episode are consistent with the linear

373 background trend defined by the VR data through this section and shallower units (Fig. 5b).

15 Holford et al. Post-Caledonian exhumation of northwest Scotland

374 Similarly, the range of palaeotemperatures indicated by AFTA data in Jurassic sample

375 GC523-58 is also consistent with this trend and palaeotemperatures indicated by VR data at

376 similar depths. Therefore we interpret the “background” linear increase of palaeotemperature

377 with depth indicated by the shallower and deeper VR data as defining a mid-Cenozoic

378 palaeothermal peak (Fig. 3) Palaeotemperatures characterising this episode are consistent

379 with a range of palaeogeothermal gradients from ~30-50 °C km-1 (Fig. 5c). If the mid-

380 Cenozoic palaeogeothermal gradient was similar to the revised present-day gradient (43°C

381 km-1), these data imply deeper burial by between 0.64 to 0.91 km of additional post-Lower

382 Jurassic section, removed during mid-Cenozoic (45-20 Ma) exhumation (Fig. 5c). These

383 results confirm km-scale deeper burial of preserved Jurassic and Triassic strata, but also

384 suggest that early Palaeocene cooling in some parts of this region, often attributed solely to

385 exhumation, is dominantly a consequence of igneous activity.

386

387 Skye, Mull and Morvern

388 We analysed 31 onshore and two offshore samples collected on and around Skye, Mull and

389 the Morvern peninsula, and one additional sample collected from central Scotland. Samples

390 encompass a broad range of ages, from Mesoproterozoic to Upper Cretaceous (Fig. 3).

391 Thermal history interpretation indicates that most samples around Skye, irrespective of their

392 stratigraphic age (e.g. Moine sample GC523-6, near Glenancross and Jurassic sample

393 GC523-10, from BH71/10) were totally annealed during early Palaeogene heating, reaching

394 temperatures >110°C before cooling which began between 65 and 60 Ma (Fig. 3). Samples

395 that were totally annealed during the early Palaeogene enable resolution of a late Cenozoic

396 cooling episode, also observed in samples from Applecross that were not totally annealed in

397 the Palaeocene event. Combining timing constraints for all samples suggests that cooling

398 began between 20 and 10 Ma (Fig. 3). Results from around Mull are generally consistent with

16 Holford et al. Post-Caledonian exhumation of northwest Scotland

399 those from Skye, although the onset of Cenozoic cooling in Devonian sample GC523-77

400 from Iona and Moine sample GC523-78 from near Bunessan is slightly later (60 to 20 Ma)

401 (Fig. 3). This could represent the failure to resolve two distinct Cenozoic cooling episodes.

402

403 AFTA results from Morvern are reported in more detail in a subsequent section of this paper,

404 but generally define distinct cooling episodes beginning during the early-mid Triassic (245 to

405 225 Ma), early-mid Cenozoic (65 to 30 Ma) and late Cenozoic (20 to 10 Ma) (Table 2).

406

407 Integrating AFTA results with regional geological evidence

408 Triassic-early Jurassic cooling episodes (245-225 and 225-190 Ma)

409 Pre-Mesozoic samples across northwest Scotland provide evidence for two phases of cooling

410 beginning during the Triassic-Jurassic. Lewisian basement samples from the Outer Hebrides

411 (GC523-95, 97, 98) and Hebrides Shelf (GC523-55) reveal Triassic-early Jurassic cooling

412 from >105°C that began between 225 and 190 Ma (Figs. 3, 6, 7a). There is no strong

413 stratigraphic evidence for major Triassic-early Jurassic erosion in the Hebridean basins (Fyfe

414 et al., 1993). Morton (1987) suggested that the first phase of crustal stretching responsible for

415 the Hebridean basin system began ~210 Ma. Because these samples are located in the

416 footwalls of major Triassic-Jurassic normal faults (e.g. the Minch Fault; Roberts &

417 Holdsworth, 1999) we interpret this cooling as due to localized exhumation driven by

418 footwall uplift. Similar footwall uplift has been identified in the North Sea from seismic data

419 (Roberts & Yielding, 1991), although any palaeothermal effects there have been overprinted

420 by Jurassic burial, reflecting the greater degree of extension and post-rift subsidence.

421

422 Mesoproterozoic (Moine), Silurian (Strontian Granite) and Devonian sedimentary AFTA

423 samples from around Skye, Mull and Morvern reveal evidence for early Triassic cooling that

17 Holford et al. Post-Caledonian exhumation of northwest Scotland

424 began between 245 and 225 Ma (Figs. 3, 6, 7a). Sample RD17-41 also shows that cooling at

425 this time also occurred in central Scotland, emphasising the regional extent of this episode

426 (Fig. 7a). Whilst the timing of cooling is similar to that observed in the footwall of the Minch

427 Fault (e.g. samples GC523-95, 97, 98), the consistency of data and interpretations over much

428 of the study area (Tables 1 and 2, Fig. 3) suggests a more regional event. One possible

429 explanation for this early Triassic cooling is that it represents the final stage of post-

430 Caledonian unroofing that brought Precambrian and Lower Palaeozoic basement rocks to the

431 surface prior to their reburial beneath Mesozoic sediments.

432

433 Early-mid Cretaceous cooling episodes (140-130 and 110-90 Ma)

434 A key finding of this study is that Permian-Lower Cretaceous sediments now exposed at or

435 near surface levels around northwest Scotland have been heated to temperatures higher than

436 90-100°C which we interpret as reflecting regional Triassic-Jurassic burial. Particularly in

437 northern and central parts of the study area (e.g. the WOB), these maximum post-depositional

438 temperatures were attained during the early to mid-Cretaceous (Figs. 3, 7b, c).

439

440 Our results indicate two separate phases of early Cretaceous cooling, with the more recent

441 event apparently restricted to the WOB (including well 202/19-1) while the earlier event

442 affected a wider region, across the Hebrides and eastwards into central Scotland (Fig. 7b).

443 Our preferred interpretation is that these are distinct episodes based on consistencies in the

444 data that characterise the different events (Tables 1 and 2). Early Cretaceous cooling is less

445 well defined around Skye, Mull and Morvern (Fig. 7b), which we attribute to overprinting by

446 early Palaeogene heating related to igneous activity. Results from well 202/19-1 suggests that

447 heating to the early to mid-Cretaceous palaeotemperatures was dominantly due to deeper

18 Holford et al. Post-Caledonian exhumation of northwest Scotland

448 burial, with subsequent cooling to exhumation, and this conclusion can be extended to

449 cooling at this time recorded in WOB shallow borehole samples (Table 2; Fig. 3).

450

451 Cretaceous rocks are largely absent around northwest Scotland, but where preserved their

452 stratigraphic relations often provide evidence for early-mid Cretaceous exhumation and

453 deformation. Cenomanian sandstones at Morvern rest unconformably on Lower Jurassic and

454 older rocks, indicating intervening uplift and erosion (Fyfe et al, 1993). In north Skye, NE-

455 trending anticlines within Jurassic strata are not observed in the thin Upper Cretaceous

456 sequences that underlie Paleocene lavas (England, 1994). The axes of these folds trend

457 parallel to the major basin-bounding faults (e.g. the Camasunary-Skerryvore Fault), and

458 indicate early Cretaceous deformation in the Hebridean region (Bell & Williamson, 2002).

459 Further south, AFTA data from wells in the Irish Sea basin system and from samples onshore

460 Ireland record widespread early Cretaceous (~120-115 Ma) exhumation (Holford et al.,

461 2009a). AFTA and compaction data from the Lower Jurassic succession of the Mochras

462 borehole indicate ~2.5 km of early Cretaceous exhumation at that location (Holford et al.,

463 2005). These observations confirm the regional nature of early Cretaceous cooling and

464 exhumation, although slight differences in the timing of cooling between individual areas

465 results in some uncertainty with regards to identification of discrete episodes and their

466 correlations with events in the stratigraphic record.

467

468 Early Cenozoic (65-60 Ma)

469 Early Cenozoic exhumation of northern Scotland is thought to be a source of Palaeocene

470 sediments in the Faroe- Basin (White & Lovell, 1997; Stoker et al., 2010), and a

471 number of previous studies have suggested regional Palaeocene uplift and exhumation across

472 northern Scotland based either on thermochronology (Persano et al., 2007) or geophysical

19 Holford et al. Post-Caledonian exhumation of northwest Scotland

473 studies (Jones et al., 2002). However, our AFTA data shows that early Cenozoic cooling is

474 restricted dominantly to the vicinity of Palaeogene igneous centres (e.g. Skye, Mull) (Fig.

475 7c). Away from these intrusions, early Cenozoic palaeotemperatures decrease rapidly, whilst

476 some regions such as the WOB show no obvious requirement for early Cenozoic cooling

477 (Fig. 7b), and earlier or later cooling episodes dominate the thermal histories.

478

479 The clearest evidence for elevated early Cenozoic palaeotemperatures is provided by Jurassic

480 AFTA samples from Skye, where all but one sample show evidence for palaeotemperatures

481 >110°C prior to cooling beginning between 65 and 60 Ma (Table 2). This is consistent with

482 previous results from Skye that also reveal early Cenozoic palaeotemperatures >110°C

483 (Lewis et al., 1992). In contrast, VR data from onshore Middle Jurassic sediments in the

484 Inner Hebrides signify localized elevated palaeotemperatures around contemporaneous

485 igneous intrusions, while away from intrusive bodies VR values are much lower (Hudson and

486 Andrews, 1987). The compaction and cementation state of these sandstones suggest

487 maximum burial depths of only ~1 km (Hudson and Andrews, 1987), indicating that the early

488 Cenozoic palaeotemperatures of >110°C revealed by AFTA most likely do not reflect solely

489 deeper burial, but are more likely due to hydrothermal and/or contact heating, similar to that

490 observed in the 134/5-1 well (Fig. 5c) Thus, whilst early Cenozoic cooling recorded by

491 AFTA data in some samples may in part record regional exhumation, over much of this

492 region it is difficult to isolate this signal from thermal effects related to igneous activity. But

493 in other areas (e.g. WOB) early Cenozoic exhumation is clearly of lower magnitude than that

494 during the early Cretaceous and mid-late Cenozoic.

495

496

497

20 Holford et al. Post-Caledonian exhumation of northwest Scotland

498 Mid-Cenozoic (45-20 Ma)

499 Data from 134/5-1 and 202/19-1 reveal evidence for mid-Cenozoic cooling. Samples from

500 134/5-1 provide the most detailed definition of this episode, showing cooling beginning

501 between 45 and 20 Ma and indicating exhumation of Lower Jurassic and older rocks

502 involving removal of ~0.6-0.9 of additional section (Fig. 5). Results from 202/19-1 are

503 consistent with a similar timing, while requiring a larger amount of removed section. This

504 cooling and exhumation is contemporaneous with a regional Upper Eocene (~34±3 Ma)

505 unconformity reported from the Atlantic margin (Praeg et al., 2005) and coeval with

506 progradation of early Oligocene clastic wedges into the eastern North Sea, which is taken to

507 record the onset of Scandinavian uplift (Japsen et al., 2007). Jolivet (2007) provides evidence

508 for similarly timed (40-25 Ma) cooling onshore Scotland recorded by apatite fission track

509 data from Loch Ness and Strontian. Jolivet (2007) interpreted this data in terms of ~1.6 to 2

510 km of post mid-Cenozoic denudation.

511

512 There are at least three outliers of late Oligocene terrestrial sediments within our study area

513 (Fig. 1). Sediments reach a maximum thickness of ~1 km in the Canna Basin, and in all areas

514 are variably tilted and folded, and overlain with marked unconformity by Quaternary deposits

515 (Fyfe et al., 1993). Lithologies (freshwater mudstones, sandstones and lignites) are similar to

516 other Oligocene terrestrial deposits in western Britain (e.g. Lough Neagh, Cardigan Bay

517 basins). Compaction and palaeothermal data show that these sediments have been exhumed

518 by up to 1.5 km (Holford et al., 2005, 2009b). No such data exist for the Oligocene outliers

519 offshore Scotland, but their preservation and deformation point to complex local variations in

520 mid-Cenozoic vertical motions.

521

522

21 Holford et al. Post-Caledonian exhumation of northwest Scotland

523 Late Cenozoic (15-10 Ma)

524 Most AFTA samples reveal major late Cenozoic cooling from palaeotemperatures of ~60-

525 65°C to their present-day temperatures (Fig. 7d; Table 1). Data from most regions are

526 consistent with a regionally synchronous onset between 15 and 10 Ma (Fig. 3, 6). AFTA

527 samples from the WOB suggest an earlier (30-15 Ma) onset, although this could represent the

528 combined effects of mid and late Cenozoic cooling. In contrast to earlier uplift episodes

529 where there is often uncertainty as to the amount of subsequent reburial, we can quantify the

530 amount of reburial following late Cenozoic uplift because Quaternary sedimentary

531 thicknesses are well constrained (generally <100 m) (e.g. Fyfe et al., 1993; Stoker et al.,

532 1993). We use late Cenozoic palaeotemperatures to estimate amounts of exhumation since

533 the late Cenozoic. Gross exhumation estimates (i.e. uncorrected for reburial) have been

534 calculated using minimum, maximum and mid-point palaeotemperature estimates, assuming

535 a late Cenozoic palaeosurface temperature of 10°C and late Cenozoic palaeogeothermal

536 gradients that decrease linearly from 43°C km-1 at 134/5-1 to 25.3°C km-1 at 202/19-1 (i.e.

537 similar to the present-day gradients) (Fig. 8). Estimates based on mid-point

538 palaeotemperatures show a consistent northwards increase, with lowest values (>0.92 km) in

539 the vicinity of Skye, and highest values (<1.94 km) in the WOB (Fig. 8). Estimates from the

540 WOB are similar to late Cenozoic exhumation estimates across the southern British Isles

541 from AFTA and sonic velocity data (Holford et al., 2005, 2009a). Around major igneous

542 centres (e.g. Mull), a significant fraction of the section eroded during the late Cenozoic may

543 have comprised lavas that accumulated during the Palaeocene (cf. Walker, 1971).

544

545 Integration of AFTA data with stratigraphic constraints on the Morvern peninsula

546 Fourteen AFTA samples have been analysed from Morvern (Fig. 9), where sedimentary rocks

547 of Mesoproterozoic (Moine metasediments; samples GC523-2, RD17-43), Upper

22 Holford et al. Post-Caledonian exhumation of northwest Scotland

548 Carboniferous (GC523-72 & 74), Triassic (GC523-1 & 3), Lower Jurassic and Upper

549 Cretaceous (Cenomanian-?Santonian; GC523-4 & 5) ages crop out within an area of ~20

550 km2. In addition there is a ~460 m thick succession of the Palaeocene Mull Lava Group,

551 whilst the eastern part of the peninsula comprises Silurian Strontian granite (RD17-42, 44 to

552 48). This area provides a unique opportunity to integrate stratigraphic constraints with

553 thermal history information from AFTA to evaluate burial and uplift histories of basement

554 and cover rocks. In basement terrains without younger sedimentary cover, the only available

555 thermal history constraints for modelling of fission track data are the measured present-day

556 temperature and an assumed depositional/metamorphic/crystallisation temperature. However,

557 Moine rocks on Morvern are variably overlain by Triassic, Lower Jurassic and Upper

558 Cretaceous sediments, showing that the basement rocks now exposed across this region had

559 been exhumed to near-surface levels/temperatures at these times.

560

561 The Upper Triassic sandstones from which samples GC523-1 and 3 were collected

562 unconformably overlie Moine Supergroup metasediments, while the Silurian Strontian

563 Granite outcrops nearby (Fig. 9a). These pre-Caledonian rocks must therefore have been

564 exhumed to surface levels by the late Triassic, at which time a regional planation surface had

565 been created, now represented by the sub-Triassic unconformity. Samples of Moine and

566 Strontian Granite from Morvern consistently record cooling from >100°C that began between

567 245 and 225 Ma (Table 2). In other words, samples now just below the level of the

568 unconformity were at paleotemperatures >100°C a few Myr prior to the time at which they

569 reached that position and the overlying Upper Triassic sedimentary units were deposited.

570 Therefore we interpret the early Triassic cooling episode as representing the period of

571 exhumation that led to formation of the unconformity underlying the Upper Triassic

572 sediments (Fig. 9c).

23 Holford et al. Post-Caledonian exhumation of northwest Scotland

573

574 The Upper Triassic sandstones are conformably overlain by Lower Jurassic sandstones and

575 shales, in turn unconformably overlain by the ~20 m thick Cenomanian (possibly Turonian)

576 marginal marine Loch Aline White Sandstone Formation, indicating the presence of a

577 Cenomanian coastline (Lowden et al., 1992). The absence of Middle Jurassic to Early

578 Cretaceous section indicates a period of erosion or non-deposition during this interval. AFTA

579 data from Morvern provide no direct evidence for cooling associated with this unconformity

580 (Fig. 9c). But as Jurassic heating and burial and early Cretaceous cooling, exhumation and

581 deformation is observed in northern parts of our study area we suggest it is also likely to have

582 occurred in Morvern, but palaeotemperatures associated with this episode were subsequently

583 overprinted by higher early Cenozoic paleotemperatures (below).

584

585 Most samples of Moine and Strontian Granite, as well as the two samples of Upper Triassic

586 sandstone, record reheating to ~80-90°C or above in the early Cenozoic (Table 2). This event

587 is also recorded in samples GC523-4 and 5 from the Loch Aline Sandstone (Table 2), which

588 show that following deposition at ~95 Ma, it was heated to 90-105°C in the Early Cenozoic.

589 Over 100 m of Upper Carboniferous (Westphalian) sandstones and mudstones occur in a

590 faulted inlier on the southern coast of Morvern. While the relationship between this and

591 younger units is not clear, thermal histories of AFTA samples GC523-72 and 74 are similar

592 to the Cenomanian samples, defining early Cenozoic palaeotemperatures of ~100°C (Fig. 9c).

593

594 As noted earlier, early Cenozoic paleotemperatures across the study region are interpreted as

595 dominantly reflecting processes related to igneous activity, chiefly contact and hydrothermal

596 effects as well as locally elevated heat flow and possibly deeper burial, and the effects of

597 these competing processes cannot be resolved. But if the early Cenozoic paleotemperatures

24 Holford et al. Post-Caledonian exhumation of northwest Scotland

598 in these samples involved any degree of deeper burial, it is clear that the entire section must

599 have been reburied following deposition of the Cenomanian sandstone, to achieve these early

600 Cenozoic palaeotemperatures. Burial beneath a late Cretaceous Chalk cover seems plausible.

601 While Upper Cretaceous strata are poorly represented in this area, sediments of this age are

602 identified in Skye, Raasay and Eigg, and Chalk Group fragments of possible Campanian age

603 occur in volcanic vents in Arran (Fyfe et al., 1993). If such burial occurred, early Palaeocene

604 exhumation must have occurred to produce the predominantly subaerial landsurface onto

605 which lavas and pyroclastic material were erupted (Emeleus & Bell, 2005). Alternatively,

606 early Cenozoic palaeotemperatures could be related to hydrothermal and other effects

607 associated with volcanic activity during or subsequent to extrusion of lavas, although in this

608 case, reheating from near-surface temperatures is still required (Fig. 9c).

609

610 Late Cenozoic (20-10 Ma) paleotemperatures around 60-70°C are also recorded in almost all

611 samples from Morvern, as well as other areas in this study (although results in some samples

612 from Morvern may represent the unresolved effects of both early and late Cenozoic episodes;

613 Table 2). The presence of the volcanic sequence and the underlying sub-aerial surface shows

614 that the section must have been re-buried. The Mull Lava Group has a thickness of ~460 m

615 on Morvern, and zeolite minerals (e.g. mesolite, laumonite) within the lavas indicate this area

616 may have been buried by >1 km of additional lavas (Bell & Williamson, 2002). The depth-

617 related zeolite zones are superimposed on hydrothermal mineral zones related to the Mull

618 Central Complex (Bell & Williamson, 2002), and a zone of Carbonate minerals on western

619 Morvern is interpreted to have been deposited by circulating heated waters from the

620 Ardnamurchan Central Complex (Walker, 1971).

621

25 Holford et al. Post-Caledonian exhumation of northwest Scotland

622 Integration of AFTA data from various stratigraphic levels with geological evidence thus

623 reveal a complex history of post-Caledonian vertical motions across Morvern (Figure 9c).

624 Moine basement rocks have been repeatedly buried then exhumed to surface levels, and at

625 least three major cycles of heating and cooling have been identified. This is in stark contrast

626 to the monotonic post-Caledonian cooling histories proposed for this region by

627 thermochronological studies that have focused on basement rocks rather than younger

628 sedimentary cover (e.g. Persano et al., 2007). We suggest that such episodic histories are

629 unlikely to be restricted to this one location, and that such histories are likely to be the rule,

630 rather than the exception.

631

632 Discussion

633 New AFTA data from northwest Scotland reveal a complex history of vertical motions

634 characterized by multiple cycles of Mesozoic-Cenozoic burial and exhumation (Fig. 10). Our

635 data reveal regional cooling and uplift episodes that began during the early-mid Triassic

636 (245-225 Ma), early Cretaceous (140-120 Ma), early Cenozoic (65-60 Ma), mid Cenozoic

637 (45-20 Ma) and late Cenozoic (15-10 Ma) (Fig. 3, 6). AFTA data in Permian-Mesozoic units

638 from offshore wells allow evaluation of palaeogeothermal gradients and indicate that deeper

639 burial is the primary cause of heating, and that exhumation is the main cause of cooling in

640 these episodes, except for early Cenozoic heating which appears in large part to be related to

641 igneous activity. The Cenozoic history of uplift and erosion is consistent with the offshore

642 sedimentary record to the north and west of our study area, with multiple tectonically-

643 influenced sedimentation pulses of Palaeocene, Eocene, Oligocene, Miocene and Plio-

644 Pleistocene observed in the Rockall and Faroe-Shetland basins (Stoker et al, 2010). Our

645 results, which indicate a major role for Cenozoic tectonics in the development of the Scottish

646 landscape, are worth comparing with recent geomorphological studies of Ireland (Dewey,

26 Holford et al. Post-Caledonian exhumation of northwest Scotland

647 2000), which indicate that many topographic features may owe their origin due to faults

648 active during Palaeogene extension and Neogene compression.

649

650 Jurassic sedimentary cover over the Highlands?

651 A key aspect of our results is that many Permian-Triassic samples appear to have reached

652 temperatures of 90-100°C or above in the early Cretaceous. Samples of basement and

653 Palaeozoic sediments give comparable results, implying similar geological histories for

654 basement terrains and sedimentary basins over the past ~200 Myr. The sedimentary basins of

655 the Hebrides contain thick successions of marine Jurassic rocks (e.g. ~2.5 km in the NMB;

656 Fyfe et al., 1993), and our results imply that comparable successions were deposited where

657 Jurassic rocks are now missing, including across the northern Highlands. Sequence-

658 stratigraphic analyses of Middle-Upper Jurassic marine sediments in the Inner Firth

659 and Hebridean basins reveal stratigraphic affinities that imply a marine connection (Underhill

660 and Partington, 1993). Underhill (1991) postulated that such a connection may have existed

661 along the trace of the Great Glen Fault. Roberts & Holdsworth (1999) argued that the

662 Highlands were traversed by a component of Mesozoic extension along NNE-SSW trending

663 faults, and that many structures that are traditionally held to be Caledonian may in fact be of

664 Mesozoic age. Quartz-filled fractures within Cambrian rocks west of the Moine Thrust Zone

665 provide independent evidence for Mesozoic extension (Laubach & Diaz-Tushman, 2009).

666 These fractures strike N-NE and are associated with post-Caledonian faults, leading Laubach

667 & Diaz-Tushman (2009) to attribute their formation to Mesozoic extension. Fluid inclusion

668 homogenization temperatures show these fractures formed at ~80°C, in good agreement with

669 early Mesozoic palaeotemperatures predicted by thermal modelling of AFTA samples that

670 from the northwest coast (Fig. 7b). AFTA indicate that a maximum of 2-3 km of Jurassic

671 sediments have been removed from northern parts of the Hebridean basin system (e.g. the

27 Holford et al. Post-Caledonian exhumation of northwest Scotland

672 West Orkney Basin) during post-Jurassic exhumation. Lower exhumation is likely in

673 southern areas where thicker Mesozoic sequences are preserved (e.g. in the vicinity of Skye,

674 where a thick Jurassic succession is preserved, diagenetic evidence suggests deeper burial of

675 these rocks by ~1 km (Hudson & Andrews, 1987)

676

677 Implications for low-temperature thermal histories of basement regions

678 The repeated cycles of heating and cooling revealed by our data and implied by the

679 geological record are in marked contrast to the results of a recent study that presented apatite

680 fission-track and (U-Th)/He (AHe) data from two locations in northwest Scotland; Clisham

681 (Achaean gneisses) and Sgorr Dhonuill (Silurian granites) (Persano et al., 2007). Persano et

682 al. (2007) interpreted their data as recording continuous, monotonic cooling over the past 400

683 My (i.e. consistent with continuous post-Caledonian emergence), with an early Cenozoic

684 acceleration in cooling. Their approach was based on forward modelling of fission-track and

685 AHe data to indentify thermal histories that were consistent with the observed fission track

686 and AHe ages.

687

688 Persano et al (2007) extracted thermal history constraints from their AHe data using the

689 systematics of He diffusion in apatite defined for Durango apatite by Farley (2000). Although

690 these systematics have been widely adopted, recent studies have shown that He diffusion in

691 apatite is more complicated than originally supposed. Green at al. (2006) reported systematic

692 inconsistencies between measured AHe ages and values predicted on the basis of thermal

693 histories derived from AFTA using the Farley (2000) systematics, with discrepancies

694 increasing with U content of the apatite and the fission track age. Further empirical data

695 showing similar effects (Flowers et al., 2007), combined with new experimental

696 investigations of AHe diffusion kinetics (Shuster et al., 2006) have shown that apatite

28 Holford et al. Post-Caledonian exhumation of northwest Scotland

697 becomes more retentive of He as radiation damage accumulates within the crystal lattice.

698 Consequently, apatites with high effective U (eU) concentrations and older fission track ages

699 retain He at much higher temperatures than young apatites with low eU concentrations such

700 as Durango apatite (Shuster et al., 2006). Because Persano et al. (2007) did not consider the

701 effects of radiation damage, the thermal history interpretations they derive from the AHe data

702 cannot be considered reliable. For all samples where Persano et al. (2007) report both AHe

703 and fission track ages, the latter are considerably older than 100 Ma (the youngest being a

704 value of 186±6 Ma sample in SD9), and so the interpretation of the AHe data in these

705 samples by Persano et al. (2007) using the Farley (2000) systematics in terms of an early

706 Cenozoic acceleration in cooling is unlikely to be valid (Green et al., 2006).

707

708 Even if this interpretation was accurate, it is far from straightforward to attribute early

709 Cenozoic cooling solely to regional exhumation, given the results here which suggest a link

710 to contemporaneous igneous activity (Fig. 2, 5b). Whilst it remains possible that regional

711 Palaeocene exhumation may contribute to early Cenozoic cooling, further work is needed to

712 isolate this signal from cooling resulting from the cessation of hydrothermal heating, contact

713 effects and more localised exhumation caused by intrusion of central complexes (Brown et

714 al., 2009).

715

716 Low-temperature thermochronological studies of basement rocks from onshore passive

717 margins and continental interiors commonly assume monotonic cooling (e.g. Dempster &

718 Persano, 2006). However, our results, best demonstrated by samples from the Morvern

719 peninsula (Fig. 9), indicate that episodic histories dominated by repeated cycles of heating

720 and cooling (i.e. burial and exhumation) are more realistic. The common assumption of

721 monotonic cooling histories for low temperature thermochronological studies is only possible

29 Holford et al. Post-Caledonian exhumation of northwest Scotland

722 occurs because such studies traditionally focus on basement rocks and ignore the significance

723 of any sedimentary outliers that may be present (Green and Duddy, 2007). We suggest that

724 this naive approach can lead to erroneous conclusions, and is no longer tenable.

725

726 Origin of uplift episodes

727 A synthesis of AFTA data from multiple wells and outcrops in the Irish Sea basins by

728 Holford et al. (2009a) identified regional (up to ~106 km2) exhumation-driven cooling

729 episodes beginning during the early Cretaceous and the early, mid and late-Cenozoic that can

730 be correlated with the episodes observed in northwest Scotland. These results suggest greater

731 complexity than previous studies that have suggested that exhumation across the western

732 British Isles is dominantly Palaeogene in age, driven by plume-related igneous underplating

733 (Brodie & White, 1994; White & Lovell, 1997; Jones et al., 2002; Persano et al., 2007).

734 Whilst underplating may well have contributed to the localized early Palaeogene exhumation

735 around Hebridean igneous centres, underplating and dynamic uplift cannot explain the

736 Triassic and early Cretaceous exhumation of northwest Scotland, and we consider it unlikely

737 that the mid and late Cenozoic exhumation episodes constrained by our AFTA data were

738 primarily plume-driven.

739

740 Holford et al. (2009a) noted close temporal relationships between the onset of Cretaceous-

741 Cenozoic exhumation episodes across the western British Isles and plate boundary

742 reorganizations associated with the North Atlantic oceanic ridge system, and increased

743 periods of convergence and coupling between the Eurasian and African plates associated with

744 Alpine collision. These correlations suggest that increased levels of intraplate stress

745 propagated from plate boundary interactions may explain the exhumation episodes

746 constrained by AFTA data. Present-day stress data from northwest Europe indicate a

30 Holford et al. Post-Caledonian exhumation of northwest Scotland

747 northwest-southeast orientation of maximum horizontal stress in the upper crust consistent

748 with control by plate boundary forces (Goelke & Coblentz, 1996). Cenozoic compressional

749 stress magnitudes in particular are likely to have been high (perhaps ~100 MPa) due to ridge

750 push forces enhanced by anomalously high temperature asthenosphere underlying the

751 elevated northeast Atlantic spreading axis, which is above sea level at Iceland (Bott, 1993;

752 Doré et al., 2008). High magnitudes of intraplate stress can explain the large number of

753 Cretaceous-Cenozoic compressional structures in post-Palaeozoic basins of the British Isles

754 and Atlantic Margin that contain weak sedimentary infills (Hillis et al., 2008). It may also

755 explain uplift of ‘uninverted’ basement regions due to the distributed effects of relatively low

756 strain rates (<10-16 s-1) operating over timescales of >1-10 Myr. The resultant cumulative

757 crustal shortening, although relatively minor (<5-10%) may nonetheless be sufficient to

758 generate long-wavelength uplift and exhumation. Low degrees of crustal shortening have

759 been invoked to explain late Cenozoic uplift and deformation of other intraplate regions e.g.

760 the West Siberian Basin (Allen & Davies, 2007) and the margins of the Australian continent

761 (Sandiford et al., 2004).

762

763 Conclusions

764 We have constrained post-Caledonian vertical motions around northwest Scotland using 78

765 AFTA samples from onshore outcrops, and offshore shallow boreholes and deep exploration

766 wells. Samples came from both Proterozoic-Lower Palaeozoic basement and Mesozoic cover

767 sequences. Our results indicate multiple Mesozoic-Cenozoic cooling and exhumation

768 episodes, several of which correlate with similar events recorded by AFTA data throughout

769 western Britain. Thermal histories of basement and cover samples are generally consistent,

770 and many Permian-Triassic samples cooled from temperatures of 90-100°C or above during

771 the Cretaceous. These results require deposition of considerable additional thicknesses of

31 Holford et al. Post-Caledonian exhumation of northwest Scotland

772 Triassic-Jurassic rocks in the Hebridean basins, and suggest deposition of Mesozoic

773 sediments across the Highlands. This additional cover was removed during regional early

774 Cretaceous and mid-late Cenozoic exhumation. AFTA data record early Cenozoic heating

775 and cooling, but it is difficult to isolate the contribution by regional uplift from thermal

776 effects associated with igneous activity. Combined with preserved stratigraphy which records

777 intermittent Upper Palaeozoic-Mesozoic sedimentation across northwest Scotland, our data

778 are incompatible with a significant Caledonian component in the present-day topography.

779 Instead, tectonic processes have repeatedly uplifted northwest Scotland over the past ~140

780 Myr and erosion has stripped away a considerable thickness of cover. Our results emphasize

781 the importance of integrating thermal history data from basement and overlying sedimentary

782 units when reconstructing the uplift histories of onshore passive margins and intraplate

783 regions, and also demonstrate the importance of integrating observations from onshore and

784 offshore.

785

786 Acknowledgements

787 This work forms part of ARC Discovery Project DP0879612, and represents TRaX

788 contribution #75. We thank Alan Roberts and an anonymous reviewer for their comments and

789 Tim Needham for editorial guidance. The contribution of Stoker is made with the permission

790 of the Executive Director, British Geological Survey (Natural Environment Research

791 Council).

792

793 References

794 ALLEN, M.B. & DAVIES, C.E. 2007. Unstable Asia: active deformation of Siberia revealed

795 by drainage shifts. Basin Research, 19, 379-392.

796

32 Holford et al. Post-Caledonian exhumation of northwest Scotland

797 BELL, B.R. & WILLIAMSON, I.T. 2002. Tertiary igneous activity. In: TREWIN, N.H. (ed)

798 The Geology of Scotland. 4th edition. Geological Society, London, 371-408.

799

800 BOTT, M.H.P. 1993. Modelling the plate-driving mechanism. Journal of the Geological

801 Society, London, 150, 941-951.

802

803 BRODIE, J. & WHITE, N. 1994. Sedimentary basin inversion caused by igneous

804 underplating. Geology, 22, 147-150.

805

806 BROWN, D.J., HOLOHAN, E.P. & BELL, B.R. 2009. Sedimentary and volcano-tectonic

807 processes in the British Paleocene Igneous Province: a review. Geological Magazine, 146,

808 326-352.

809

810 BURNHAM, A.K. & SWEENEY, J.J. 1989. A chemical kinetic model of vitrinite reflectance

811 and maturation. Geochimica et Cosmochimica Acta, 53, 2649-2657.

812

813 CHADWICK, R.A. & PHAROAH, T.C. 1998. The seismic reflection Moho beneath the

814 United Kingdom and adjacent areas. Tectonophysics, 299, 255-279.

815

816 DEMPSTER, T.J. & PERSANO, C. 2006. Low-temperature thermochronology: Resolving

817 geotherm shapes or denudation histories? Geology, 34, 73-76.

818

819 DEWEY, J.F. 2000. Cenozoic tectonics of western Ireland. Proceedings of the Geologists’

820 Association, 111, 291-306.

821

33 Holford et al. Post-Caledonian exhumation of northwest Scotland

822 DORÉ, A.G., CARTWRIGHT, J.A., STOKER, M.S., TURNER, J.P. & WHITE, N.J. 2002.

823 Exhumation of the North Atlantic margin: introduction and background. In: DORÉ, A.G.,

824 CARTWRIGHT, J.A., STOKER, M.S., TURNER, J.P. & WHITE, N. (eds) Exhumation of

825 the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration.

826 Geological Society, London, Special Publications, 196, 1-12.

827

828 DORÉ, A.G., LUNDIN, E.R., KUSZNIR, N.J. & PASCAL, C. 2008. Potential mechanisms

829 for the genesis of Cenozoic domal structures on the NE Atlantic margin: pros, cons and some

830 new ideas. In: JOHNSON, H., DORÉ, A. G., GATLIFF, R. W., HOLDSWORTH, R.,

831 LUNDIN, E. & RITCHIE, J. D. (eds) The Nature and Origin of Compression in Passive

832 Margins. Geological Society, London, Special Publications, 306, 1–26.

833

834 EMELEUS, C. H. & BELL, B. R. 2005. The Palaeogene volcanic districts of Scotland. 4th

835 ed. British Geological Survey, Nottingham.

836

837 ENGLAND, R.W. 1994. The structure of the Skye lava field. Scottish Journal of Geology,

838 30, 33-37.

839

840 EVANS, D.J. 1997. Estimates of the eroded overburden and the Permian-Quaternary

841 subsidence history of the area west of Orkney. Scottish Journal of Geology, 33, 169-182.

842

843 FARLEY, K.A. 2000. Helium diffusion from apatite: General behaviour as illustrated by

844 Durango fluorapatite. Journal of Geophysical Research, 105B, 2903-2914.

845

34 Holford et al. Post-Caledonian exhumation of northwest Scotland

846 FLOWERS, R.M., SHUSTER, D.L., WERNICKE, B.P. & FARLEY, K.A. 2007. Radiation

847 damage control on apatite (U-Th)/He dates from the Grand Canyon region, Colorado Plateau.

848 Geology, 35, 447-450.

849

850 FYFE, J. A., LONG, D. & EVANS, D. 1993. United Kingdom offshore regional report: The

851 Geology of the Malin-Hebrides Sea Area. British Geological Surey, London, Her Majesty’s

852 Stationary Office.

853

854 GALBRAITH, R.F. & LASLETT, G.M. 1993. Statistical methods for mixed fission track

855 ages. Nuclear Tracks, 21, 459-470.

856

857 GALLAGHER, K. 1995. Evolving temperature histories from apatite fission-track data.

858 Earth and Planetary Science Letters, 136, 421-435.

859

860 GEIKIE, A. 1901. The Scenery of Scotland. London.

861

862 GEORGE, T.N. 1966. Geomorphic evolution in Hebridean Scotland. Scottish Journal of

863 Geology, 2, 1-34.

864

865 GÖLKE, M. & COBLENTZ, D. 1996. Origins of the European regional stress field.

866 Tectonophysics, 266, 11-24.

867

868 GREEN, P.F. 1986. On the thermo-tectonic evolution of Northern England: evidence from

869 fission track analysis. Geological Magazine, 123, 493-506.

870

35 Holford et al. Post-Caledonian exhumation of northwest Scotland

871 GREEN, P.F. & DUDDY, I.R., 2007. Discussion on ‘Apatite (U – Th)/He age constraints on

872 the Mesozoic and Cenozoic evolution of the Bathurst Region, New South Wales: evidence

873 for antiquity of the continental drainage divide along a passive margin’. Australian Journal of

874 Earth Sciences, 54, 1009–1011.

875

876 GREEN, P.F., THOMSON, K. & HUDSON, J.D. 2001. Recognizing tectonic events in

877 undeformed regions: contrasting results from the Midland Platform and East Midlands Shelf,

878 Central England. Journal of the Geological Society, London, 158, 59-73.

879

880 GREEN, P.F., DUDDY, I.R. & HEGARY, K.A. 2002. Quantifying exhumation in

881 sedimentary basins of the UK from apatite fission track analysis and vitrinite reflectance data:

882 precision, accuracy and latest results. In: DORÉ, A.G., CARTWRIGHT, J.A., STOKER,

883 M.S., TURNER, J.P. & WHITE, N. (eds) Exhumation of the North Atlantic Margin: Timing,

884 Mechanisms and Implications for Petroleum Exploration. Geological Society, London,

885 Special Publications, 196, 331-354.

886

887 GREEN, P.F., CROWHURST, P.V., DUDDY, I.R., JAPSEN, P. & HOLFORD, S.P. 2006.

888 Conflicting (U-Th)/He and fission track ages in apatite: Enhanced He retention, not

889 anomalous annealing behaviour. Earth and Planetary Science Letters, 250, 407-427.

890

891 HALL, A.M. 1991. Pre-Quaternary landscape evolution in the Scottish Highlands.

892 Transactions of the Royal Society of Edinburgh: Earth Sciences, 82, 1-26.

893

894 HALLAM, A. 1983. Jurassic, Cretaceous and Tertiary sediments. In: CRAIG, G.Y. (ed)

895 Geology of Scotland. 2nd edition. Scottish Academic Press, Edinburgh, 343-356.

36 Holford et al. Post-Caledonian exhumation of northwest Scotland

896 HILLIER, S. & MARSHALL, J.E.A. 1992. Organic maturation, thermal history and

897 hydrocarbon generation in the Orcadian Basin, Scotland. Journal of the Geological Society,

898 London, 149, 491-502.

899

900 HILLIS, R.R., HOLFORD, S.P., GREEN, P.F., DORÉ, A.G., GATLIFF, R.W., STOKER,

901 M.S., THOMSON, K., TURNER, J.P., UNDERHILL, J.R. & WILLIAMS, G.A. 2008.

902 Cenozoic exhumation of the southern British Isles. Geology, 36, 371-374.

903

904 HOLFORD, S.P., GREEN, P.F. & TURNER, J.P. 2005. Palaeothermal and compaction

905 studies in the Mochras borehole (NW Wales) reveal early Cretaceous and Neogene

906 exhumation and argue against regional Palaeogene uplift in the southern Irish Sea. Journal of

907 the Geological Society, London, 162, 829-840.

908

909 HOLFORD, S.P., GREEN, P.F., DUDDY, I.R., TURNER, J.P., HILLIS, R.R. & STOKER,

910 M.S. 2009a. Regional intraplate exhumation episodes related to plate boundary deformation.

911 Geological Society of America Bulletin, 121, 1611-1628, doi: 10.1130/B26481.1

912

913 HOLFORD, S.P., GREEN, P.F., HILLIS, R.R., TURNER, J.P. & STEVENSON, C.T.E.

914 2009b. Mesozoic-Cenozoic exhumation and volcanism in Northern Ireland constrained by

915 AFTA and compaction data from the Larne No. 2 borehole. Petroleum Geoscience, 15, 239-

916 257.

917

918 HUDSON, J.D. & ANDREWS, J.E. 1987. The diagenesis of the Great Estuarine Group,

919 Middle Jurassic, Inner Hebrides, Scotland. In: Marshall, J.D. (ed.) Diagenesis of Sedimentary

920 Sequences. Geological Society, London, Special Publications, 36, 259-276.

37 Holford et al. Post-Caledonian exhumation of northwest Scotland

921 HURFORD, A.J. & GREEN, P.F. 1983. The zeta age calibration of fission-track dating.

922 Chemical Geology, 1, 285-317.

923

924 JAPSEN, P. 2000. Investigation of multi-phase erosion using reconstructed shale trends

925 based on sonic data, Sole Pit axis, North Sea. Global and Planetary Change, 24, 189-210.

926

927 JAPSEN, P., BIDSTRUP, T. & LIDMAR-BERGSTRÖM, K. 2002. Neogene uplift and

928 erosion of southern Scandinavia induced by the rise of the South Swedish Dome. In: DORÉ,

929 A.G., CARTWRIGHT, J.A., STOKER, M.S., TURNER, J.P. & WHITE, N. (eds)

930 Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for

931 Petroleum Exploration. Geological Society, London, Special Publications, 196, 183-207.

932

933 JAPSEN, P., GREEN, P.F., HENRIK NIELSEN, L., RASMUSSEN, E.S. & BIDSTRUP, T.,

934 2007. Mesozoic-Cenozoic exhumation events in the eastern North Sea Basin: a multi-

935 disciplinary study based on paleothermal, palaeoburial, stratigraphic and seismic data. Basin

936 Research, 19, 451-490.

937

938 JOLIVET, M. 2007. Histoire de la denudation dans le corridor du loch Ness (Écosse):

939 mouvements verticaux différentiels le long de la Great Glen Fault. Comptes Rendus

940 Geoscience, 339, 121-131.

941

942 JONES, S.M., WHITE, N., CLARKE, B.J., ROWLEY, E. & GALLAGHER, K. 2002.

943 Present and past influence of the Iceland Plume on sedimentation. In: DORÉ, A.G.,

944 CARTWRIGHT, J.A., STOKER, M.S., TURNER, J.P. & WHITE, N. (eds) Exhumation of

38 Holford et al. Post-Caledonian exhumation of northwest Scotland

945 the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration.

946 Geological Society, London, Special Publications, 196, 13-25.

947

948 LAUBACH, S.E. & DIAZ-TUSHMAN, K. 2009. Laurentian palaeostress trajectories and

949 ephemeral fracture permeability, Cambrian Eriboll Formation sandstones west of the Moine

950 Thrust Zone, NW Scotland. Journal of the Geological Society, London, 166, 349-362.

951

952 LEWIS, C. L. E., CARTER, A. & HURFORD, A.J. 1992. Low temperature effects of the

953 Skye Tertiary intrusions on Mesozoic sediments in the Sea of Hebrides Basin. In:

954 PARNELL, J. (ed) Basins on the Atlantic Seaboard: Petroleum Geology, Sedimentology and

955 Basin Evolution. Geological Society, London, Special Publications, 62, 64-77.

956

957 LOWDEN, B., BRALEY, S., HURST, A. & LEWIS, J. 1992. Sedimentological studies of

958 the Cretaceous Lochaline Sandstone, NW Scotland. In: PARNELL, J. (ed) Basins on the

959 Atlantic Seaboard: Petroleum Geology, Sedimentology and Basin Evolution. Geological

960 Society, London, Special Publications, 62, 159-162.

961

962 MACDONALD, D.I.M., ARCHER, B., MURRAY, S., SMITH, K., BATES, A. 2007.

963 Modelling and comparing the Caledonian and Permo-Triassic erosion surfaces across

964 Highland Scotland: implications for landscape inheritance. In: NICHOLS, G., WILLIAMS,

965 E. & PAOLA, C. (eds) Sedimentary Processes, Environments and Basins: A Tribute to Peter

966 Friend. Special Publications of the IAS, 38, 283-299.

967

39 Holford et al. Post-Caledonian exhumation of northwest Scotland

968 MAGARA, K. 1976. Thickness of removed sedimentary rocks, palaeopore pressure and

969 palaeotemperature, south-western part of Western Canada Basin. AAPG Bulletin, 60, 554-

970 565.

971

972 MARIE, J.P.P. 1975. Rotliegendes stratigraphy and diagenesis. In: WOODLAND, A.W.

973 (ed.) Petroleum and the Continental Shelf of North-west Europe, Volume 1, Geology, Applied

974 Science Publishers, 205-211.

975

976 MORTON, N. 1987. Jurassic subsidence history in the Hebrides, N.W. Scotland. Marine and

977 Petroleum Geology, 4, 226-241.

978

979 NIELSEN, S.B., GALLAGHER, K., LEIGHTON, C., BALLING, N., SVENNINGSEN, L.,

980 HOLM JACOBSEN, B., THOMSEN, E., NIELSEN, O.B., HEILMANN-CLAUSEN,

981 EGHOLM, D.L., SUMMERFIELD, M.A., RØNØ CLAUSEN, O., PIOTROWSKI, J.A.,

982 THORSEN, M.R., HUUSE, M., ABRAHAMSEN, N., KING, C. & LYKKE-ANDERSEN,

983 H. 2009. The evolution of western Scandinavian topography: a review of Neogene uplift

984 versus the ICE (isostasy-climate-erosion) hypothesis. Journal of Geodynamics, 47, 72-95.

985

986 PERSANO, C., BARFORD, D.N., STUART, F.M. & BISHOP, P. 2007. Constraints on early

987 Cenozoic underplating-driven uplift and denudation of western Scotland from low

988 temperature thermochronometry. Earth and Planetary Science Letters, 263, 404-419.

989

990 PRAEG, D., STOKER, M.S., SHANNON, P.M., CERAMICOLA, S., HJELSTUN, B.,

991 LABERG, J.S. & MATHIESEN, A. 2005. Episodic Cenozoic tectonism and the development

40 Holford et al. Post-Caledonian exhumation of northwest Scotland

992 of the NW European ‘passive’ continental margin. Marine and Petroleum Geology, 22, 1007-

993 1030.

994

995 ROBERTS, A.M. & YIELDING, G. 1991. Deformation around basin-margin faults in the

996 North Sea/mid-Norway rift. In: ROBERTS, A.M., YIELDING. G. & FREEMAN, B. (eds)

997 The Geometry of Normal Faults. Geological Society, London, Special Publications, 56, 61-

998 78.

999

1000 ROBERTS, A.M. & HOLDSWORTH, R.E. 1999. Linking onshore and offshore structures:

1001 Mesozoic extension in the Scottish Highlands. Journal of the Geological Society, 156, 1061-

1002 1064.

1003

1004 SANDIFORD, M., WALLACE, M. & COBLENTZ, D. 2004. Origin of the in situ stress field

1005 in south-eastern Australia. Basin Research, 16, 325-338.

1006

1007 SHAW CHAMPION, M.E., WHITE, N.J., JONES, S.M. & LOVELL, J.P.B. 2008.

1008 Quantifying transient mantle convective uplift: an example from the Faroe-Shetland basin.

1009 Tectonics, 27, TC1002, doi:10.1029/2007TC002106.

1010

1011 SHUSTER, D.L., FLOWERS, R.M. & FARLEY, K.A. 2006. The influence of natural

1012 radiation damage on helium diffusion kinetics in apatite. Earth and Planetary Science

1013 Letters, 249, 148-161.

1014

41 Holford et al. Post-Caledonian exhumation of northwest Scotland

1015 STOKER, M.S., HITCHEN, K. & GRAHAM, C.G. 1993. United Kingdom Offshore

1016 Regional Report: the Geology of the Hebrides and West Shetland Shelves, and Adjacent

1017 Deep-Water Areas. British Geological Survey. HMSO, London.

1018

1019 STOKER, M.S., HOLFORD, S.P., HILLIS, R.R., GREEN, P.F. & DUDDY, I.R. 2010.

1020 Cenozoic post-rift sedimentation off Northwest Britain: Recording the detritus of episodic

1021 uplift on a passive continental margin. Geology, 38.

1022

1023 THOMSON, K., UNDERHILL, J.R., GREEN, P.F., BRAY, R.J. & GIBSON, H.J. 1999.

1024 Evidence from apatite fission-track analysis for the post-Devonian burial and exhumation

1025 history of the northern Highlands, Scotland. Marine and Petroleum Geology, 16, 27-39.

1026

1027 TURNER, J.P., GREEN, P.F., HOLFORD, S.P. & LAWRENCE, S. 2008. Thermal history

1028 of the obliquely divergent Rio Muni (West Africa)-NE Brazil margins during continental

1029 breakup. Earth and Planetary Science Letters, 270, 354-367.

1030

1031 UNDERHILL, J. R. 1991. Implications of Mesozoic Recent basin development in the

1032 western lnner Moray Firth, UK. Marine & Petroleum Geology, 8, 359 369.

1033

1034 UNDERHILL, J.R. 2001. Controls on the genesis and prospectivity of Paleogene

1035 palaeogeomorphic traps, East Shetland Platform, UK North Sea. Marine and Petroleum

1036 Geology, 18, 259-281.

1037

1038 UNDERHILL, J.R. & PARTINGTON, M.A. 1993. Jurassic thermal doming and deflation in

1039 the North Sea: implications of the sequence stratigraphic evidence. In: PARKER, J.R. (ed.)

42 Holford et al. Post-Caledonian exhumation of northwest Scotland

1040 Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference. Geological

1041 Society, London, 337-345.

1042

1043 UNDERHILL, J.R., MONAGHAN, A.A. & BROWNE, M.A.E. 2008. Controls on structural

1044 styles, basin development and petroleum prospectivity in the Midland Valley of Scotland.

1045 Marine and Petroleum Geology, 15, 1000-1022.

1046

1047 WALKER, G.P.L. 1971. The distribution of amygdale minerals in Mull and Morven

1048 (Western Scotland). In: MURTY, T.V.V.G.G.R.K. & ROA, S.S. (eds) Studies in Earth

1049 Sciences (W.D. West Commemoration Volume). University of Saugur, India, 181-194.

1050

1051 WATSON, J. 1985. Northern Scotland as an Atlantic-North Sea divide. Journal of the

1052 Geological Society, London, 142, 221-243.

1053

1054 WHITE, N. & LOVELL, B. 1997. Measuring the pulse of a plume with the sedimentary

1055 record. Nature, 387, 888-891.

1056

1057 Figure captions

1058 Fig. 1 Map of Scotland showing post-Ordovician geology and faults with evidence of post-Ordovician activity,

1059 and locations of samples used in this study (Fyfe et al., 1993; Stoker et al., 1993). Although Mesozoic and

1060 Cenozoic rocks are mainly restricted to the present-day offshore basins, sedimentary and volcanic outliers of

1061 Devonian, Carboniferous, Permian, Upper Triassic, Jurassic and Upper Cretaceous occur along the northwest

1062 coast. CB, Canna Basin; CF, Colonsay Fault; CL, Clisham; GGF, Great Glen Fault; HBF, Highland Boundary

1063 Fault; HF, Helmsdale Fault; HS, Hebrides Shelf; IMF, Inner Moray Firth; LMT, Little Minch Trough; M, Mull;

1064 MF, Minch Fault; NH, Northern Highlands; NMB, North Minch Basin; OH, Outer Hebrides; S, Skye; SD, Sgorr

1065 Dhonuill; SOHB, Sea of Hebrides Basin; SKF, Skerryvore Fault; STF, Strathconon Fault; WF, Wick Fault;

1066 WFB, West Flannan Basin; WLB, West Lewis Basin; WOB, West Orkney Basin; WSB, West Shetland Basin.

43 Holford et al. Post-Caledonian exhumation of northwest Scotland

1067 Fig. 2 Mean fission track ages plotted against mean track lengths for apatite fission track data used in this study,

1068 alongside previously published data from northwest Scotland (Lewis et al., 1992 (margins of Sea of Hebrides

1069 Basin); Thomson et al., 1999 (northern Highlands); Jolivet, 2007 (Loch Ness, Strontian); Persano et al., 2007

1070 (Clisham, Sgorr Dhonuill)). All errors shown at ±1σ. Most fission track ages are younger than 300 Ma,

1071 suggesting these samples reached late Palaeozoic-Mesozoic palaeotemperatures sufficient to produce severe age

1072 reduction (i.e. >100°C). Ages from Skye and Mull mostly range between 50-70 Ma, similar to the timing of

1073 igneous activity in these regions, and suggesting that related thermal effects heavily influence fission track data

1074 from these regions. A typical ‘boomerang trend’ based on AFTA data from the Lake District (Green, 1986) is

1075 shown in the background.

1076

1077 Fig. 3 Comparison of the timing information derived from all AFTA data that show evidence for higher

1078 palaeotemperatures post-deposition/intrusion/metamorphism. Synthesis of results identifies a number of discrete

1079 phases of cooling, indicated by the vertical bands. Open boxes and arrow indicate the range of stratigraphic ages

1080 for individual samples. Horizontal bands define range of timing for the onset of cooling derived from AFTA

1081 data in each sample within a 95% confidence interval. Note that for several samples from 202/19-1 the timing

1082 estimates for individual estimates show some overlap. Timing estimates have been shifted vertically where this

1083 is the case.

1084

1085 Fig. 4 AFTA results and thermal history reconstruction for exploration well 202/19-1. a) Left, fission-track

1086 ages; right, mean track lengths, both plotted against depth below kelly bushing. Black lines show increasing

1087 stratigraphic age with depth. Greyscale lines show the predicted patterns of fission-track age and mean track

1088 length from the DTH for apatites containing 0.0-0.1, 0.4-0.5, 0.9-1.0 and1.5-1.6 wt% Cl. The DTH is based on

1089 the burial history derived from the sedimentary section intersected by the well, combined with the present-day

1090 geothermal gradient of 25.3°C km-1 determined from corrected BHT values in this well. The fission-track ages

1091 decrease with depth and are much less than the values predicted from the DTH, indicating that the sampled

1092 sedimentary units have been hotter in the past. b) Palaeotemperature constraints from AFTA data plotted against

1093 depth. AFTA data define two separate phases of cooling that have affected the preserved Permian-Triassic

1094 succession that began during the early Cretaceous (110 to 90 Ma) and Cenozoic (55 to 10 Ma). c) Ranges of

1095 amounts of removed section and palaeogeothermal gradients (hyperbolic ellipsoids) required to explain the early

1096 Cretaceous and Cenozoic palaeotemperatures. The hyperbolic ellipsoids define the parameter ranges estimated

44 Holford et al. Post-Caledonian exhumation of northwest Scotland

1097 from AFTA that are consistent with the respective palaeotemperature constraints within 95% confidence limits.

1098 Black dots indicate the corresponding maximum likelihood solutions. Horizontal shaded band shows amount of

1099 removed section independently estimated from sonic velocity data (Evans, 1997).

1100

1101 Fig. 5 AFTA results and thermal history reconstruction for exploration well 134/5-1. a) Left, fission-track ages;

1102 right, mean track lengths, both plotted against depth below kelly bushing. Black lines show increasing

1103 stratigraphic age with depth. Greyscale lines show predicted patterns of fission-track age and mean track length

1104 from the DTH based on the preserved sedimentary section for apatites containing 0.0-0.1, 0.4-0.5, 0.9-1.0

1105 and1.5-1.6 wt% Cl. The DTH for this well has been determined using the revised geothermal gradient of 43°C

1106 km-1 that is with AFTA data. b) Palaeotemperature constraints from AFTA and vitrinite reflectance (VR) data

1107 plotted against depth. AFTA and VR data define two episodes of Cenozoic cooling. Samples recording early

1108 Cenozoic (~60 Ma) palaeotemperatures are closely associated with coeval igneous intrusions and define no

1109 linear trend with depth, suggesting that cooling occurred following the cessation of contact and hydrothermal

1110 heating associated with igneous activity. Samples recording mid-Cenozoic (45 to 20 Ma) palaeotemperatures

1111 define a linear profile with depth which has a similar gradient to that of the present-day temperature profile. (c)

1112 Mid-Cenozoic cooling in these samples is explained by removal of ~0.6 to 0.9 km of post-Lower Jurassic

1113 section during a denudation episode which began between 45 and 20 Ma, for a palaeogeothermal gradient equal

1114 to the revised present-day value of 43°C km-1.

1115

1116 Fig. 6 Post-400 Ma event diagram for northwest Scotland, comparing generalized stratigraphy with AFTA-

1117 constrained cooling episodes defined in this study and in the western British Isles (Holford et al., 2009a).

1118

1119 Fig. 7 Palaeotemperatures across northwest Scotland in each of the regional palaeothermal episodes defined in

1120 this study. a) Triassic (220-195 Ma in the Outer Hebrides and 245-225 Ma in Morvern and central Scotland).

1121 Lighter shading shows surface/seafloor outcrops of Permian-Triassic rocks, darker shading shows outcrops of

1122 Jurassic rocks. b) Early Cretaceous (110-90 Ma in the West Orkney Basin and 140-130 Ma in the Hebridean

1123 region). Lighter shading shows surface/seafloor outcrops of Permian-Triassic rocks, darker shading shows

1124 outcrops of Jurassic rocks. c) Early Palaeogene (65-60 Ma in Skye, Mull and Morvern, 75-40 Ma in the

1125 Hebridean region). Diagonal shading shows seafloor outcrops of Cenozoic sedimentary rocks, hatched shading

1126 shows surface/seafloor outcrops of Palaeogene igneous extrusive rocks, grey shading shows surface/seafloor

45 Holford et al. Post-Caledonian exhumation of northwest Scotland

1127 outcrops of Palaeogene intrusive rocks. d) Mid-late Cenozoic (30-15 Ma in West Orkney Basin, 15-10 Ma in

1128 Hebridean region, 40-25 Ma in 134/5-1 and 20-10 Ma in Skye, Mull and Morvern).

1129

1130 Fig. 8 Mid-late Cenozoic exhumation estimates across northwest Scotland based on AFTA data. a) Estimates

1131 based on mid-point temperature for each sample, assuming a surface temperature of 10°C and a geothermal

1132 gradient that decreases linearly from 43°C km-1 at 134/5-1 to 25.3°C km-1 at 202/19-1. b) Horizontal bars

1133 represent the range of gross exhumation for each sample location (reflecting the range of palaeotemperatures

1134 estimated from AFTA), and thickness of Quaternary deposits (i.e. post-exhumation reburial). Adding the

1135 thickness of Quaternary sediments to the gross estimates gives an estimate of net exhumation for each location.

1136 Results indicate that large areas of northwest Scotland and its offshore basins have had >0.5-1.0 km of section

1137 removed during late Cenozoic exhumation.

1138

1139 Fig. 9 AFTA data from the Morvern peninsula. a) Geological map of the Morvern peninsula. b) Thermal

1140 histories for AFTA samples from various stratigraphic levels. Results indicate exhumation of basement rocks to

1141 surface levels by the late Triassic, following which they were reheated during late-Triassic-Jurassic burial, and

1142 then cooled through several stages of Cretaceous-Cenozoic exhumation. c) Schematic post-400 Ma thermal

1143 history of the Morvern peninsula reconciling results from all AFTA samples.

1144

1145 Fig. 10 Schematic ~E-W cross-sections illustrating our preferred model for the post-Caledonian geological

1146 evolution of northwest Scotland. (a) Early Devonian post-orogenic uplift witnessed by the unroofing of the

1147 Newer Granites (Watson, 1985; Hall, 1991). (b) Devonian extensional collapse and subsidence permitting

1148 deposition of >2-3 km of Devonian and Lower Carboniferous rocks (Thomson et al., 1999). (c) Late

1149 Carboniferous deformation and uplift related to far-field Variscan compression (Underhill et al., 2008)

1150 removing Devonian and Carboniferous sediments. (d) Late Triassic-Jurassic extension and subsidence leading to

1151 formation of Hebridean basins (Fyfe et al., 1993) but also affecting pre-Caledonian basement rocks (Roberts &

1152 Holdsworth, 1999). Localised uplift in footwalls of major faults, and regional uplift of basement rocks in

1153 western and central Scotland. Maximum post-Caledonian burial depths across northwest Scotland are attained

1154 during the late Jurassic-early Cretaceous with up to 2-3 km of subsequently eroded syn and post-rift Jurassic

1155 rocks deposited across this region. (e) Early Cretaceous regional uplift and exhumation beginning ~120 Ma

1156 followed by late Cretaceous sea level rises permitting Chalk deposition (f). (g) Localised early Cenozoic uplift

46 Holford et al. Post-Caledonian exhumation of northwest Scotland

1157 (beginning ~65-60 Ma) around Palaeogene igneous intrusive centres. Considerable thicknesses of lavas

1158 accumulate in some areas (e.g. around Mull), much of which are eroded during later exhumation. (h) Regional

1159 exhumation beginning during the mid-Cenozoic (40-25 Ma) and late Cenozoic (15-0 Ma) removing up to 1.9

1160 km of overburden, followed by Pliocene-Pleistocene uplift not directly constrained by our AFTA data but

1161 witnessed by shelf-slope wedges of clastic sediment that prograde away from northwest Scotland (Stoker et al.,

1162 1993).

1163

1164 Table 1 Apatite fission track analysis sample details.

1165

1166 Table 2 Summary of palaeotemperatures analysis and Default Thermal History interpretations.

47 Holford et al Figure 1

Carboniferous cz Cenozoic sediments j c pt Jurassic sediments sediments Permian-Triassic g Palaeogene intrusives pt d Devonian extrusives sediments Carboniferous d g Palaeogene extrusives c d Devonian sediments extrusives WSB k Cretaceous sediments k, j, pt g 202/19-1 WLB 503-151 SBHF cz d pt 503-146 pt

523-24 503-147 523-47 503-148 pt g WOB g HS MF j 503-149 pt 503-150 WF 523-51 d 523-52 523-48 g 503-184 k, j, pt 523-15 IMF WFB NH cz 523-49 503-185 NMB k 58°N 523-23 OH 523-110 HF 523-98 523-13 CL 523-111 g 523-93 523-14 523-97 523-95 523-112 LMT 523-11 j 523-91 pt g 523-99 523-102 523-105 d g 523-109 S 523-106 STF j 523-55 523-108 g 523-83, 84 523-80 CB 523-88 RD17-41 cz 523-89 j pt 523-10 GGF 134/5-1 523-6,8 523-56 MF pt RD17-47,48 RD17-42 SOHB RD17-45,46 SD d pt 523-1,2,3 RD17-44 g 523-4,5 RD17-43 523-72,74 523-79 Fig. 9 523-71 HBF SKF 523-77 M d d cz 523-78 d

g 56°N 523-9

8°W pt CF 6°Wc 4°W c 2°W Holford et al Figure 2

16 West Lewis/West Orkney Basin Hebrides 15 Skye, Mull and Morvern Lewis et al. (1992) Thomson et al. (1999) 14 Jolivet (2007) Persano et al. (2007) Lake District ‘boomerang trend’ 13 Green (1986)

12

11

10

9

8 0 50 100 150 200 250 300 350 400 450 500 Fission track age (Ma) Holford et al Figure 3 Time (Ma) Time (Ma) 300 250 200 150 100 50 0 300 250 200 150 100 50 0

Permian Triassic Jurassic Cretaceous Cenozoic Permian Triassic Jurassic Cretaceous Cenozoic

West Orkney Basin 134/5-1 GC369-1 GC523-24 GC369-2 GC503-146 GC523-68 GC503-147 GC369-3 GC503-148 GC369-4 GC503-149 GC369-5 GC503-150 Degree of overlap 45-20 Ma GC503-151 Skye, Mull and Morvern 202/19-1 GC503-84 GC523-1 GC503-85 GC523-2 >800 GC503-86 GC523-3 GC503-87 GC523-4 GC503-88 GC523-5 GC503-89 GC523-6 >800 GC503-90 GC523-8 >800 GC503-91 GC523-9 416-359 GC503-92 GC523-10 416 Degree of overlap 110-90 Ma 30-15 Ma GC523-71 -385 GC523-72 345-328 Hebrides GC523-74 345-328 GC503-184 >570 GC523-77 416-359 GC503-185 GC523-78 >800 GC523-11 GC523-79 GC523-13 1100-800 GC523-83 GC523-14 GC523-84 GC523-15 GC523-88 1100-800 GC523-23 GC523-89 GC523-47 GC523-91 GC523-48 GC523-99 GC523-49 GC523-102 GC523-51 GC523-105 1100-800 GC523-52 GC523-106 GC523-55 2900-1100 GC523-108 GC523-56 1100 GC523-109 -800 GC523-80 >800 RD17-41 >800 GC523-93 2900-1100 RD17-42 420 GC523-95 2900-1100 RD17-43 >800 GC523-97 2900-1100 RD17-44 420 GC523-98 2900-1100 RD17-45 420 GC523-110 2900-1100 RD17-46 420 GC523-111 1100-800 RD17-47 420 GC523-112 1900-1100 RD17-48 420 225-190 Ma 140-130 Ma 75-40 Ma 15-10 Ma 245-225 Ma 135-100 Ma 65-60 Ma 20-10 Ma

300 250 200 150 100 50 0 300 250 200 150 100 50 0 Time (Ma) Time (Ma) Holford et al Figure 4 (a) 0 Quat 10 10

20 20 1 1 30 30

40 40

50 50 2 2

Permian-Triassic 60 60 Depth (km)

70 (°C) Temperature 70 3 3 80 80 stratigraphic age

4 0 100 200 300 0 4 8 12 16 Fission Track Age (Ma) Mean Length (µm) (b) Temperature (°C) 20 40 60 80 100 120 140 160 180 200 Sea bed 127 mKB Quat

GC503-84 0.5

GC503-85

1.0 GC503-86

1.5 GC503-87

GC503-88 GC503-89

Depth (km) 2.0 GC503-90 Upper Permian-Lower Triassic Upper Permian-Lower

GC503-91 2.5

GC503-92 3.0

TD 3189 m Corrected BHT measurement AFTA sampling horizon Maximum mid-late Cenozoic (55-10 Ma) palaeotemperature from AFTA Maximum early Cretaceous (110-90 Ma) palaeotemperature from AFTA (c)

10 Early Cretaceous maximum likelihood estimates Palaeogeothermal gradient: 39°C/km Additional section: 2.31 km 8 Cenozoic maximum likelihood estimates Palaeogeothermal gradient: 18.2°C/km Additional section: 2.67 km

6 Present-day geothermal gradient = 25.3°C/km

4 3.9 km

Additional section (km) 3.3 km

2 Triassic shale sonic velocities >1.85 km 1.7 km 1.2 km

20 40 60 80 100 Palaeogeothermal gradient° ( C/km) Holford et al Figure 5 (a) 0 10 10 20 20 30 30 1 1 40 40 stratigraphic age 50 50 Lower Jurassic 60 60 70 70 Temperature (°C) Temperature Depth (km) 2 2 80 80 90 90 Triassic

3 0 100 200 300 400 0 4 8 12 16 Fission T rack Age (Ma) Mean Length (µm) (b) T emperature (°C) 20 40 60 80 100 120 140 160 180 200 220 240 260 Sea bed 132 mKB

Mid-Cenozoic Pliensbachian 0.5 (45 to 20 Ma) palaeogeothermal gradient GC523-58 Inferred early Cenozoic palaeotemperature profile 1.0 GC369-1 Late Sinemurian

Early 1.5 Sinemurian

Depth (km) Present-day GC369-2 geothermal Hettangian gradient of Rhaetian GC523-68 ~43°C/km Palaeogene 2.0 GC369-3 (based on BHT intrusive rocks

and AFTA data) GC369-4 GC369-5 Triassic

2.5 TD 2472 m Corrected BHT measurement AFTA sampling horizon Revised present-day temperature derived from AFTA data Maximum mid-Cenozoic (45 to 20 Ma) palaeotemperature from AFTA Maximum mid-Cenozoic (45 to 20 Ma) palaeotemperature from VR Maximum early Cenozoic palaeotemperature from AFTA Maximum early Cenozoic palaeotemperature from VR (c) Mid-late Cenozoic maximum likelihood estimates 2 Palaeogeothermal gradient: 39.4°C/km Additional section: 0.94 km

1.5

Present-day geothermal gradient = 43°C/km 1 0.91 km

Additional section (km) 0.64 km 0.5

20 30 40 50 60 70 Palaeogeothermal gradient (°C/km) Holford et al Figure 6

Age Generalized AF TA-constrained (Ma) stratigraphy cooling episodes 0 Widespread 15-10 Ma 20-10 Ma 20-15 Ma 30-15 Ma Canna Basin 40-25 Ma

Widespread 75-40 Ma 65-60 Ma 65-55 Ma

Morvern, Skye, Mull, 100 Arran 110-90 Ma

Early Cretaceous 120-115 Ma

Cretaceousdeformation Cenozoic and uplift 140-130 Ma

Callovian hiatus Toarcian hiatus Widespread

200 Upper Triassic rocks overly 225-190 Ma basement in Morvern 245-225 Ma

Base poorly dated Variscan deformation and uplift 300 Morvern

Arran 134/5-1 Hebridean Region West Orkney Basin West

Mull, British Isles Western Arran Skye, Mull & Morvern 400 Devonian Carboniferous Permian Triassic Jurassic Holford et al Figure 7 a) b)

58°N 58°N

Triassic Early Cretaceous palaeotemperatures palaeotemperatures >110°C >110°C 100-110°C 100-110°C 90-100°C 90-100°C 80-90°C 80-90°C 70-80°C 70-80°C 60-70°C 60-70°C <60°C <60°C

56°N 56°N

8°W 6°W 4°W 2°W 8°W 6°W 4°W 2°W

8°W 6°W 4°W 2°W c) d)

58°N 58°N

Early Palaeogene Mid-Late Cenozoic palaeotemperatures palaeotemperatures >110°C >110°C 100-110°C 100-110°C 90-100°C 90-100°C 80-90°C 80-90°C 70-80°C 70-80°C 60-70°C 60-70°C <60°C <60°C

56°N 56°N

8°W 6°W 4°W 2°W 8°W 6°W 4°W 2°W Holford et al Figure 8 Late Cenozoic exhumation (km) b) 0 0.5 1.0 1.5 2.0 2.5 3.0

1.67 GC503-151 84 m 0.74 2.59 a) 8°W 6°W 4°W 2°W 1.92 GC503-146 23 m 1.54 2.30 1.69 GC523-47 52 m 1.45 1.93 1.94 GC503-148 8.5 m 1.29 2.59 1.94 GC503-147 28 m 1.29 2.59 1.72 GC503-149 28.5 m 0.86 2.59 1.43 GC503-150 12.5 m 0.71 2.14 1.47 GC523-51 8.5 m 1.18 1.76 1.62 GC523-48 12.1 m 1.32 1.91 1.47 GC523-52 10.5 m 1.03 1.91 1.32 GC523-49 10.6 m 0.97 1.67 1.74 GC523-23 1.85 m 1.39 2.08 1.64 GC523-11 9 m 1.28 2.00 1.15 GC523-91 0 m 0.54 1.76 1.01 GC523-99 0 m 0.54 1.49 1.01 GC523-102 0 m 0.54 1.49 1.49 GC523-105 0 m 1.22 1.76 1.64 GC523-109 0 m 1.18 2.10 1.10 58°N GC523-55 68 m 0.93 1.28 1.47 RD17-41 0 m 1.2 1.73 0.92 GC523-106 0 m 0.39 1.45 1.35 GC523-83 0 m 0.64 2.05 1.35 GC523-84 0 m 0.90 1.79 1.11 GC523-88 0 m 0.53 1.71 Late Cenozoic 1.22 GC523-89 0 0.26 2.18 exhumation 1.52 GC523-10 19 m 1.34 1.71 1.8-2.0 km 1.13 GC523-6 0 m 0.88 1.38 1.44 1.6-1.8 km RD17-42 0 m 1.2 1.69 1.55 1.4-1.6 km RD17-47 0 m 1.43 1.67 0.88 1.2-1.4 km RD17-48 0 m 0.59 1.18 1.55 1.0-1.2 km RD17-45 0 m 1.43 1.67 1.55 0.8-1.0 km RD17-46 0 m 1.43 1.67 1.56 RD17-44 0 m 1.44 1.68 1.08 RD17-43 0 m 0.72 1.44 0.88 GC523-4 0 m 0.47 1.28 1.33 GC523-5 0 m 0.93 1.74 1.22 GC523-72 0 m 0.81 1.63 1.22 GC523-74 0 m 0.47 1.98 1.34 56°N GC523-79 0 m 1.16 1.51 1.34 GC523-78 0 m 0.81 1.86 0500 Quaternary reburial (m) Holford et al Figure 9

Time (Ma) Time (Ma) 800 400 300 200 100 0 350300250 200 150 100 50 0 Paleocene lavas b) 0 0 Upper Cretaceous RD17-42,45,46,47,48 RD17-44 GC523-2 Moine (Meso- Lower Jurassic 25 proterozoic) 25 Upper Triassic Upper Carboniferous RD17-43 50 50 Strontian Granite (Silurian) Moine (Mesoproterozoic) Fault GC523-1,2 75 75 Temperature (°C) Temperature Temperature (°C) Temperature

GC523-3 100 100 GC523-72 and GC523-74 Carboniferous GC523-4 125 125 Time (Ma) Time (Ma) 450400 350 300 250 200 150 100 50 0 10080 60 40 20 0 0 0 GC523-4 and GC523-5 RD17-44 Upper Cretaceous GC523-5 Strontian 25 Granite 25 Sound of Mull (Silurian) Temperature (°C) Temperature

(Underlain by Moine) 50 50 Loch Morvern ? Alineline

GC523-4 75 75 Temperature (°C) Temperature Temperature (°C) Temperature

GC523-72 100 100 GC523-74

a) 125 125 0 km 4

Time (Ma) c) 400 350 300250 200 150 100 50 0 0 Summary of events Upper Upper Upper Palaeocene Unroofing of Strontian Granite prior Carboniferous Triassic Cretaceous basalts 1 to Carboniferous (and Devonian?) sandstone sandstone sandstone sedimentation 2 Late Carboniferous-Permian burial 25 and heating Moine and Strontian samples from 3 Morvern consistent with cooling 5 7 from >100°C between 245-225 Ma 1 10 4 Early-mid Triassic cooling and 2 8 exhumation 50 5 Triassic-Jurassic burial and heating Early Cretaceous cooling not 4 required by these data (overprinted 6 by Palaeocene heating?) but exhumation likely based on regional 11 results 75 Late Cretaceous burial and heating Temperature (°C) Temperature 6 7 (beneath eroded Chalk cover) Latest Cretaceous uplift to produce 8 subaerial landsurface onto which lavas were erupted Early Cenozoic palaeotemperatures 100 of 90°C or above due to either burial 9 9 beneath a lava cover of hydrothermal and contact heating associated with 3 igneous intrusions 10 Burial by Cenozoic sediments? 125 65- 40-25, 11 Mid-late Cenozoic cooling from Regional cooling episodes 245-225 Ma 140-130, 110-90 Ma 60 Ma 20-15 Ma ~65-70°C due to regional exhumation Holford et al Figure 10

(a) Early Devonian - uplift (b) Mid Devonian-Carboniferous - subsidence Carboniferous sediments onlapping and covering parts of the Highlands?

Rapidly unroofed Newer Granites Devonian ORS Lower Palaeozoic- e.g. Outer Isles e.g. Moine Precambrian basement Thrust Thrust Extensional reactivation of Caledonian thrusts cf. DRUM seismic profile

(c) Late Carboniferous - uplift (d) Triassic-Jurassic - basin subsidence, basement uplift Folding, inversion and erosion Uplifting footwall Early Triassic uplift of Carboniferous sediments e.g. Outer Hebrides Jurassic onshore Scotland post-rift Sea-level

Devonian ORS Triassic syn-rift Basement (including e.g. Minch Devonian ORS) Fault

(e) Early Cretaceous - uplift (f) Late Cretaceous - subsidence Upper Cretaceous Chalk deposited Regional uplift and erosion occurring Jurassic over sedimentary basins and basement in multiple stages and leading to post-rift Sea-level removal of Jurassic sediments

Triassic Cretaceous Jurassic syn-rift Triassic post-rift post-rift syn-rift

Continuing extension to the northwest

(g) Early Cenozoic - magmatism and localised uplift (h) Mid-Late Cenozoic - uplift Quaternary glaciation and rejuvenation Sills and dykes intruded Central complex Erosion and removal Fissure-fed lava fields of Chalk and lava of topography into Mesozoic sediments e.g. Skye e.g. Mull Sea-level

Reactivation of Hot fluid circulation basin bounding Uplift of basins and faults? basement

Localized uplift associated with igneous activity Sample location Uranium Fission track age P [χ2] (number Mean track S.D. (μm) (number Sample number Stratigraphic age (Ma) (depth bsl if offshore) content (ppm) (Ma ± 1σ)a of grains) length (μm) of track lengths) West Orkney Basin GC523-24 BGS rockdrill 58-06/642 (4.7 m) Permian-Triassic (299-200) 4 76.5 ± 13.7 17 (20) 12.12 ± 0.60 2.55 (18) GC503-146 BGS BH72/37 (9-21 m) Triassic (251-200) 17 239.3 ± 22.8 <1 (19) 11.90 ± 0.19 1.87 (100) GC503-147 BGS BH73/31 (28-31 m) Permian-Triassic (299-200) 9 300.4 ± 23.9 <1 (20) 10.98 ± 0.20 2.04 (100) GC503-148 BGS BH72/34 (14 m) Lower Jurassic (200-176) 12 201.5 ± 32.1 <1 (12) 11.26 ± 0.35 2.48 (51) GC503-149 BGS BH73/29 (31-36 m) Permian-Triassic (299-200) 13 281.5 ± 15.2 14 (20) 11.50 ± 0.17 1.83 (118) GC503-150 BGS BH72/25 (15-19m) Permian-Triassic (299-200) 10 295.8 ± 23.7 <1 (20) 11.88 ± 0.17 1.74 (101) GC503-151 BGS BH78/07 (125-128 m) Permian-Triassic (299-200) 10 428.6 ± 33.0 <1 (20) 11.54 ± 0.18 1.84 (109)

202/19-1 GC503-84 274-427 m Upper Permian (260-251) 22 138.6 ± 17.7 <1 (20) 12.27 ± 0.23 2.00 (75) GC503-85 610-762 m Upper Permian (260-251) 13 162.5 ± 19.3 <1 (20) 13.21 ± 0.19 1.77 (84) GC503-86 975-1128 m Upper Permian (260-251) 15 138.1 ± 17.0 <1 (20) 12.30 ± 0.28 2.11 (57) GC503-87 1372-1524 m Upper Permian (260-251) 10 82.5 ± 9.1 6 (20) 11.88 ± 0.51 1.52 (9) GC503-88 1753-1758 m Upper Permian (260-251) 28 65.8 ± 4.9 5 (20) 12.99 ± 0.16 1.23 (56) GC503-89 1810-1819 m Upper Permian (260-251) 23 107.8 ± 16.6 <1 (20) 12.81 ± 0.33 1.58 (23) GC503-90 1999-2008 m Upper Permian (260-251) 17 70.1 ± 8.5 93 (20) 11.82 ± 0.60 1.88 (10) GC503-91 2316-2469 m Upper Permian (260-251) 26 58.5 ± 7.2 <1 (20) 12.03 ± 0.31 1.60 (27) GC503-92 2743-2895 m Upper Permian (260-251) 21 57.4 ± 6.0 <1 (20) 11.08 ± 0.62 2.24 (13)

Hebrides GC503-184 Balchrick Torridonian (1000-530) 24 176.4 ± 15.7 <1 (20) 12.41 ± 0.24 2.36 (100) GC503-185 Tolm, Lewis Permian-Triassic (299-200) 20 212.1 ± 16.0 <1 (20) 12.11 ± 0.28 2.10 (58) GC523-11 BGS BH71/11 (19-25 m) Permian-Triassic (299-200) 12 119.2 ± 13.2 <1 (20) 10.88 ± 0.29 2.65 (82) GC523-13 BGS BH71/13 (14.5-15 m) Torridonian (1000-530) 15 133.2 ± 32.0 <1 (15) 10.08 ± 0.54 2.21 (17) GC523-14 BGS BH71/14 (23-27 m) Permian-Triassic (299-200) 11 72.2 ± 18.8 <1 (13) 11.05 ± 0.95 3.16 (11) GC523-15 BGS BH72/32 (28-36 m) Permian-Triassic (299-200) 8 139.5 ± 23.0 4 (13) 12.08 ± 0.52 2.50 (23) GC523-23 BGS rockdrill 58-08/94 (1.85-2.6 m) Permian-Triassic (299-200) 17 160.1 ± 15.0 <1 (19) 11.63 ± 0.26 2.68 (106) GC523-47 BGS BH78/05 (54 m) Permian-Triassic (299-200) 7 187.7 ± 17.2 13 (20) 12.68 ± 0.19 1.93 (104) GC523-48 BGS BH88/04 (11.15-12.61 m) Permian-Triassic (299-200) 11 184.3 ± 22.7 <1 (20) 12.37 ± 0.22 2.30 (106) GC523-49 BGS BH88/08 (11.5-12.1 m) Permian-Triassic (299-200) 13 219.8 ± 20.2 3 (20) 11.45 ± 0.27 2.75 (106) GC523-51 BGS BH90/08 (17.25-12 m) Upper Jurassic (161-145) 27 270.3 ± 27.5 <1 (20) 12.11 ± 0.17 1.91 (125) GC523-52 BGS BH90/09 (18-22.3 m) Lower Cretaceous (145-100) 7 220.8 ± 19.0 71 (20) 12.35 ± 0.26 1.52 (34) GC523-55 BGS BH90/14 (91 m) Lewisian (>2500) 18 238.9 ± 22.1 <1 (20) 12.89 ± 0.15 1.65 (114) GC523-56 BGS BH90/16 (75-76.5 m) Permian-Triassic (299-200) 9 215.7 ± 16.9 7 (20) 11.80 ± 0.31 2.40 (62) GC523-80 Dornie, Loch Duich Moine (1100-800) 15 227.3 ± 15.2 21 (20) 11.39 ± 0.26 2.68 (106) GC523-93 Luskentyre, Harris Lewisian (>2500) 4 70.8 ± 9.2 17 (20) 14.25 ± 0.69 1.70 (6) GC523-95 Northton, Harris Lewisian (>2500) 11 201.4 ± 11.3 22 (20) 14.40 ± 0.15 1.48 (102) GC523-97 Meavag, Harris Lewisian (>2500) 4 180.9 ± 15.3 70 (20) 12.95 ± 0.36 2.13 (35) GC523-98 Bowglass, Harris Lewisian (>2500) 6 216.7 ± 15.2 17 (20) 13.91 ± 0.12 1.26 (111) GC523-110 Achmelvich Bay Lewisian (>2500) 2 147.7 ± 26.8 75 (20) 11.79 ± 0.35 2.37 (46) GC523-111 Loch Lurgainn Torridonian (1000-530) 18 296.1 ± 33.6 <1 (7) 12.15 ± 0.25 2.20 (77) GC523-112 Gruinard Bay Lewisian (>2500) 1 234.2 ± 76.7 89 (20) 12.13 ± 0.85 1.90 (5)

134/5-1 GC523-58 579-731 m Pliensbachian (190-183) 13 256.7 ± 41.3 9 (5) 12.11 ± 0.42 0.84 (4) GC369-1 930-1082 m Sinemurian (197-190) 26 81.1 ± 37.6 <1 (5) 10.11 ± 1.88 4.20 (5) GC369-2 1524-1676 m Hettangian-Sinemurian (200-190) 16 57.7 ± 33.0 <1 (4) 14.30 ± 0.27 0.38 (2) GC523-68 1829-1966 m Triassic (251-200) 9 74.5 ± 37.5 <1 (8) 12.46 ± 0.45 1.09 (6) GC369-3 1966-2079 m Triassic (251-200) 16 19.6 ± 4.9 <1 (19) 10.30 ± 0.35 1.88 (29) GC369-4 2164-2255 m Triassic (251-200) 33 8.7 ± 2.1 <1 (22) 11.24 ± 0.47 1.04 (5) GC369-5 2255-2347 m Triassic (251-200) 32 7.6 ± 0.8 14 (20) 10.54 ± 0.79 3.06 (15)

Skye, Mull and Morvern GC523-1 Loch Aline, Morvern Triassic (251-200) 18 229.2 ± 13.6 63 (20) 10.66 ± 0.21 2.13 (102) GC523-2 Loch Aline, Morvern Moine (1100-800) 18 206.4 ± 14.3 52 (20) 10.43 ± 0.21 2.25 (110) GC523-3 Loch Aline, Morvern Triassic (251-200) 18 208.1 ± 12.3 69 (20) 10.75 ± 0.21 2.17 (107) GC523-4 Loch Aline, Morvern Cenomanian (100-94) 21 221.2 ± 14.5 79 (20) 10.99 ± 0.22 2.18 (102) GC523-5 Loch Aline, Morvern Cenomanian (100-94) 15 219.2 ± 18.2 <1 (17) 10.82 ± 0.37 2.19 (35) GC523-6 Glenancross Moine (1100-800) 8 59.5 ± 7.6 98 (20) 11.43 ± 0.27 2.70 (103) GC523-8 Glenancross Moine (1100-800) 15 177.8 ± 11.5 21 (20) 10.82 ± 0.23 2.29 (103) GC523-9 BGS BH71/09 (42-46 m) Devonian (416-359) 5 231.1 ± 23.0 23 (20) 10.87 ± 0.27 2.01 (57) GC523-10 BGS BH71/10 (20-22 m) Jurassic (200-145) 26 60.6 ± 15.7 <1 (14) 13.04 ± 0.17 1.83 (115) GC523-71 Ganavan Bay Lower-Middle Devonian (416-385) 11 228.4 ± 26.5 <1 (19) 12.08 ± 0.24 2.30 (88) GC523-72 Inninmore Bay, Morvern Visean (345-328) 13 157.0 ± 11.0 33 (20) 10.83 ± 0.42 2.80 (44) GC523-74 Inninmore Bay, Morvern Visean (345-328) 10 142.6 ± 14.5 <1 (20) 10.53 ± 0.33 3.44 (108) GC523-77 Fionnphort, Mull Devonian (416-359) 7 203.7 ± 16.9 9 (20) 11.48 ± 0.22 2.20 (101) GC523-78 Bunessan, Mull Moine (1100-800) 24 151.6 ± 9.1 28 (20) 10.12 ± 0.26 2.75 (110) GC523-79 Balnahard, Mull Triassic (251-200) 11 67.9 ± 5.5 20 (20) 12.58 ± 0.22 2.27 (102) GC523-83 Broadford Bay, Skye Lower Jurassic (200-176) 21 52.3 ± 6.0 <1 (20) 13.53 ± 0.29 1.66 (32) GC523-84 Broadford Bay, Skye Lower Jurassic (200-176) 16 54.7 ± 4.7 3 (20) 13.91 ± 0.24 1.53 (40) GC523-88 Drocaid Lusa, Skye Torridonian (1000-530) 23 61.9 ± 4.4 19 (20) 13.43 ± 0.41 2.20 (29) GC523-89 Loch Slapin, Skye Middle Jurassic (176-161) 12 71.2 ± 8.2 <1 (20) 14.39 ± 0.23 1.55 (45) GC523-91 Cairidh Ghlumaig, Skye Middle Jurassic (176-161) 13 219.2 ± 18.4 <1 (20) 12.11 ± 0.30 2.53 (72) GC523-99 Staffin, Skye Middle Jurassic (176-161) 22 66.4 ± 4.9 9 (20) 13.92 ± 0.27 0.99 (13) GC523-102 Culnacnoc, Skye Middle Jurassic (176-161) 19 68.0 ± 6.5 <1 (20) 14.14 ± 0.18 1.12 (39) GC523-105 Cuaig Torridonian (1000-530) 26 60.6 ± 15.7 <1 (14) 12.60 ± 0.32 2.44 (60) GC523-106 Applecross Lower Jurassic (200-176) 13 275.7 ± 18.9 8 (19) 10.85 ± 0.20 2.11 (108) GC523-108 Applecross Permian-Triassic (299-200) 16 179.6 ± 34.6 <1 (10) 10.19 ± 0.43 2.21 (27) GC523-109 Applecross Torridonian (1000-530) 13 270.9 ± 22.2 4 (20) 10.35 ± 0.26 2.69 (108) RD17-41 Fort Augustus Moine (1100-800) 29 179.8 ±10.6 <1 (20) 11.98 ± 0.14 1.41 (104) RD17-42 Strontian (A884) Strontian Granite (420) 37 207.9 ± 8.5 96 (20) 11.76 ± 0.13 1.43 (120) RD17-43 Coire nam Muc (A884) Moine (1100-800) 36 213.7 ± 9.1 23 (20) 11.89 ± 0.14 1.48 (109) RD17-44 Beinn Chlaonleud (A884) Strontian Granite (420) 32 187.7 ± 9.2 10 (20) 11.96 ± 0.11 1.13 (100) RD17-45 Beathrach (A884) Strontian Granite (420) 32 195.1 ± 13.0 <1 (20) 12.02 ± 0.11 1.08 (102) RD17-46 Beathrach (A884) Strontian Granite (420) 32 204.1 ± 12.3 <1 (20) 12.14 ± 0.12 1.28 (108) RD17-47 Liddesdale (A884) Strontian Granite (420) 30 202.9 ± 12.6 <1 (20) 12.28 ± 0.13 1.27 (101) RD17-48 Liddesdale (A884) Strontian Granite (420) 28 188.6 ± 11.1 <1 (20) 12.02 ± 0.14 1.40 (100)

Table 1. Apatite fission track analysis sample details aCentral age (Galbraith and Laslett, 1993) used for samples containing a significant spread in single grain ages (P[χ2]>5%), otherwise the “pooled age” is quoted. All ages were calculated using the zeta calibration approach of Hurford and Green (1983) using zeta values of 336.0±6.0 for samples GC369-1 to 5 (SRM612 dosimeter glass), 380.4±5.7 for samples GC503-84 to 92 (CN5 dosimeter glass), 354.8±6.3 for GC503-184 and 185 (SRM612), 386.9±6.9 for GC523-88 to 102 (CN5), 378.4±5.5 for GC503-146 to 151 (CN5), 392.9±7.4 for GC523-1 to 84 and GC523-105-112 (CN5), 348.8±5.2 for RD17-41 and 44-48 (SRM612) and 360.3±6.8 for RD17-42 and 43 (SRM612). All errors quoted at ±1σ. All analytical details are as described by Green et al. (2001), with the exception that the thermal neutron irradiation showed a significant flux gradient, and the appropriate value of ρD was determined by linear interpolation through the stack of grain mounts.

Predicted FTA Predicted Cooling episodes from DTH MTL from Pre-Cretaceous Cretaceous Early Cenozoic Mid-Late Cenozoic Sample (Ma); DTH (μm); Max. Max. Max. Max. number Onset of Onset of Onset of Onset of measured FTA measured palaeotemperature palaeotemperature palaeotemperature palaeotemperature cooling (Ma) cooling (Ma) cooling (Ma) cooling (Ma) (Ma) MTL (μm) (°C) (°C) (°C) (°C) West Orkney

Basin GC523-24 208-290 (76.5) 14.7 (12.12) ≥105 200-75 90-100 65-15 GC503-146 146-210 (239.3) 14.7 (11.90) 85-95 190-185 50-70 50-0 GC503-147 208-290 (300.4) 14.7 (10.98) 85-95 180-130 45-80 30-0 GC503-148 178-208 (201.5) 14.7 (11.26) 90-100 185-55 45-80 55-0 GC503-149 208-290 (281.5) 14.7 (11.50) 85-110 ≥45 35-85 55-0 GC503-150 208-290 (295.8) 14.7 (11.88) 80-90 190-45 30-70 45-0 GC503-151 208-290 (428.6) 14.7 (11.54) 80-105 ≥50 30-80 70-0

202/19-1 GC503-84 246 (138.6) 14.7 (12.27) 95-100 155-80 40-65 55-0 GC503-85 246 (162.5) 14.4 (13.21) ≥100 180-70 50-90 120-0 GC503-86 246 (138.1) 14.1 (12.30) ≥100 150-70 50-90 90-0 GC503-87 243 (82.5) 13.7 (11.88) ≥110 200-90 GC503-88 238 (65.8) 13.2 (12.99) ≥110 130-70 GC503-89 239 (107.8) 13.2 (12.81) ≥120 130-80 60-105 105-0 GC503-90 236 (70.1) 12.8 (11.82) ≥120 130-70 60-110 80-0 GC503-91 225 (58.5) 12.0 (12.03) ≥110 110-50 80-105 55-0 GC503-92 200 (57.4) 10.8 (11.08) ≥115 180-70 90-110 65-10 Overlap 110-90 Ma 30-15 Ma

Hebrides GC503-184 570 (176.4) 14.7 (12.41) 100-105 180-110 75-85 80-40 GC503-185 208-290 (212.1) 14.7 (12.11) 75-90 155-30 GC523-11 208-290 (119.2) 14.7 (10.88) 95-100 75-30 55-80 15-0 GC523-13 800-1100 (133.2) 14.7 (10.08) 115-125 270-90 90-100 80-0 GC523-14 208-290 (72.2) 14.7 (11.05) 100-105 90-10 GC523-15 208-290 (139.5) 14.7 (12.08) 105-115 200-100 80-95 90-20 GC523-23 208-290 (160.1) 14.7 (11.63) 95-105 140-75 60-85 50-0 GC523-47 208-245 (187.7) 14.7 (12.68) 90-105 205-130 55-70 70-5 GC523-48 208-245 (184.3) 14.7 (12.37) 95-105 180-95 55-75 55-0 GC523-49 208-245 (219.8) 14.7 (11.45) 95-100 155-75 45-70 30-0 GC523-51 208-290 (270.3) 14.7 (12.11) 80-95 260-100 50-70 40-0 GC523-52 208-290 (220.8) 14.7 (14.51) 80-105 270-80 45-75 70-0 GC523-55 1100-2900 (238.9) 14.7 (12.89) 95-105 270-170 50-65 50-0 GC523-56 208-245 (215.7) 14.7 (11.80) 85-95 170-70 45-65 25-0 GC523-80 800 (227.3) 14.7 (11.39) 100-105 235-120 70-80 80-10 GC523-93 1100-2900 (70.8) 14.7 (14.25) 100-110 100-40 GC523-95 1100-2900 (201.4) 14.7 (14.40) ≥120 225-175 45-70 100-55 GC523-97 1100-2900 (180.9) 14.7 (12.95) 105-110 240-140 45-75 100-0 GC523-98 1100-2900 (216.7) 14.7 (13.91) ≥105 270-190 45-65 120-20 GC523-110 1100-2900 (147.7) 14.7 (11.79) 100-110 250-70 75-90 65-5 GC523-111 800-1100 (296.1) 14.7 (12.15) 90-100 270-100 50-75 80-5 GC523-112 1100-1900 (234.2) 14.7 (12.13) 70-100 240-0 70-100 240-0 70-100 240-0 Overlap 225-190 Ma 140-130 Ma 75-40 Ma 15-10 Ma

134/5-1 GC523-58 184 (256.7) 14.0 (12.11) 45-90 195-0 GC369-1 180 (81.1) 13.5 (10.11) 95-100 125-0 GC369-2 167 (57.7) 12.7 (14.30) ≥125 85-25 GC523-68 129 (74.5) 9.9 (12.46) ≥100 245-20 GC369-3 96 (19.6) 10.3 (10.30) ≥120 50-20 GC369-4 29 (8.7) 10.0 (11.24) ≥115 65-10 GC369-5 18 (7.6) 10.6 (10.54) ≥125 45-10 Overlap 85-25 Ma 45-20 Ma

Skye, Mull

and Morvern GC523-1 208-290 (229.2) 14.7 (10.66) 85-90 70-10 GC523-2 800 (206.4) 14.7 (10.43) 100-120 370-100 80-95 70-10 GC523-3 208-290 (208.1) 14.7 (10.75) 80-90 65-10 GC523-4 90-97 (221.2) 14.7 (10.99) 90-105 90-25 30-65 30-0 GC523-5 90-97 (219.2) 14.7 (10.82) 90-110 97-30 50-85 30-0 GC523-6 800 (59.5) 14.7 (11.43) ≥100 70-50 45-65 25-0 GC523-8 800 (177.8) 14.7 (10.82) 100-105 255-100 85-95 30-10 GC523-9 363-403 (231.1) 14.7 (10.87) ≥100 400-210 85-95 95-5 GC523-10 146-208 (60.6) 14.7 (13.04) ≥125 70-50 65-80 25-10 GC523-71 386-408 (228.4) 14.7 (12.08) 110-120 285-190 80-90 90-35 GC523-72 333-350 (157.0) 14.7 (10.83) 95-100 100-25 45-80 25-0 GC523-74 333-350 (142.6) 14.7 (10.53) 100 70-20 30-95 25-0 GC523-77 363-408 (203.7) 14.7 (11.48) 95-105 305-135 80-90 60-20 GC523-78 800 (151.6) 14.7 (10.12) 95-100 60-20 45-90 20-0 GC523-79 208-245 (67.9) 14.7 (12.58) 100-105 75-35 60-75 25-5 GC523-83 178-208 (52.3) 14.7 (13.53) ≥115 65-40 35-90 40-0 GC523-84 178-208 (54.7) 14.7 (13.91) ≥120 70-50 45-80 45-10 GC523-88 800-1100 (61.9) 14.7 (13.43) ≥110 75-55 30-75 50-0 GC523-89 157-178 (71.2) 14.7(14.39) ≥120 95-60 20-95 65-0 GC523-91 157-178 (219.2) 14.7 (12.11) 85-95 135-55 30-75 80-10 GC523-99 157-178 (66.4) 14.7 (13.92) ≥105 85-50 30-65 45-0 GC523-102 157-178 (68.0) 14.7 (14.14) ≥110 80-55 30-65 45-0 GC523-105 800-1100 (60.6) 14.7 (12.60) 115-125 90-60 55-75 40-0 GC523-106 178-208 (275.7) 14.7 (10.85) 85-90 80-20 25-65 20-0 GC523-108 208-290 (179.6) 14.7 (10.19) 90-100 65-0 GC523-109 800-1100 (270.9) 14.7 (10.35) 95-100 280-50 55-90 40-0 RD17-41 800 (179.8) 14.7 (11.98) >100 245-195 80-90 175-60 55-75 30-0 RD17-42 420 (207.9) 14.7 (11.76) >100 280-225 75-95 225-25 60-80 25-0 RD17-43 800 (213.7) 14.7 (11.89) >100 280-170 75-90 175-25 40-70 25-0 RD17-44 420 (187.7) 14.7 (11.96) >100 260-205 70-85 205-30 70-80 30-0 RD17-45 420 (195.1) 14.7 (12.02) >100 270-225 70-80 30-0 RD17-46 420 (204.1) 14.7 (12.14) >100 265-225 70-80 35-0 RD17-47 420 (202.9) 14.7 (12.28) >100 270-220 70-80 45-0 RD17-48 420 (188.6) 14.7 (12.02) >100 255-200 80-85 100-25 35-65 20-0 Overlap 245-225 Ma 175-100 Ma 65-60 Ma 20-10 Ma

Table 2. Summary of palaeotemperatures analysis and Default Thermal History interpretation