<<

1 Age of the Sääksjärvi , Finland: reconciling the

2 timing of small impacts in crystalline basement with regional

3 basin development

4

5 Gavin G. Kenny1*, Irmeli Mänttäri2, Martin Schmieder3,4, Martin J. Whitehouse1,

6 Alexander A. Nemchin1,5, Jeremy J. Bellucci1, Renaud E. Merle1

7

8 1Department of Geosciences, Swedish Museum of Natural History, SE-104 05 Stockholm,

9

10 2Radiation and Nuclear Safety Authority, Laippatie 4, Box 14, FI-00811 Helsinki, Finland

11 3Lunar and Planetary Institute – USRA, Houston TX 77058, USA

12 4HNU – Neu-Ulm University of Applied Sciences, D-89231 Neu-Ulm, Germany

13 5School of and Planetary Sciences (EPS), Curtin University, Perth, WA 6845, Australia

14

15 *[email protected]

16

17 ABSTRACT

18 We report a new age for the Sääksjärvi impact structure, Finland, a 6 km in diameter

19 feature that formed in crystalline rocks of the Precambrian Baltic Shield. Two previous studies

20 have reported 40Ar/39Ar data for Sääksjärvi and suggested conflicting formation ages of ≤330 Ma

21 or approximately 560 Ma, respectively. The former represents a possible complication for

22 models which indicate that the region was covered by sediments of the Caledonian foreland

23 basin throughout much of the Phanerozoic. We conducted a study combining imaging,

24 microstructural analysis and U–Pb dating of shocked zircon from Sääksjärvi. The U–Pb dataset

25 indicates a ca. 600 Ma impact into predominantly ca. 1850 Ma target rocks. A concordia age of

26 608 ± 8 Ma (2σ) confirms Sääksjärvi as the first known Ediacaran impact structure in the Baltic

27 Shield and only the second worldwide. Our data indicate that the Sääksjärvi impact structure

28 formed in exposed crystalline basement rocks of the Baltic Shield prior to the development of the

29 Caledonian foreland basin. Given that most impact structures on Earth are relatively small

30 features, radiometric dating of small impact structures in crystalline basement may place

31 boundaries on the timing and spatial extent of palaeobasins that might be otherwise difficult to

32 constrain.

33

34 1. INTRODUCTION

35 Small impact structures in crystalline basement have a unique ability to shed light on the

36 temporal and spatial extent of palaeobasins. Impact structures sometimes provide, in their impact

37 lithologies or down-faulted segments, a useful record of pre-impact bedrock units that have

38 otherwise been completely removed from around the structure; for example, the Lappajärvi

39 impact structure, Finland, preserves in its structural depression sedimentary units that have

40 facilitated reconstructions of Cambrian palaeogeography in the Baltic region (e.g., Slater and

41 Willman, 2019, and references therein). Similarly, the impact structure in Sweden, and the

42 Manicouagan impact structure in Canada, also preserve megablocks of early Phanerozoic

43 sedimentary units that are not well preserved in the surrounding regions (e.g., Grieve, 2006;

44 Grieve and Head, 1983; Lehnert et al., 2012, 2014, and references therein). At Paasselkä, and a

45 number of other impact structures in Finland and Sweden, impact melt contain shocked

46 and melted fragments of Mesoproterozoic ‘ Sandstone’ that has been completely removed

47 by erosion elsewhere in the area (Buchner et al., 2009; Puura and Plado, 2005). Conversely, the

48 presence of small impact structures in crystalline basement can provide evidence that the

49 basement was not deeply covered by sediments at a given time if an accurate and precise age for

50 the structure can be attained. Such structures may therefore offer an opportunity to test, and even

51 place boundaries on, the size and lifetime of sedimentary basins which might otherwise be

52 difficult to constrain. To explore this possibility, we here revisit the age of a small impact

53 structure in crystalline basement (the Sääksjärvi impact structure, Finland) for which conflicting

54 40Ar/39Ar ages have been published (Bottomley et al., 1990; Müller et al., 1990), one of which

55 overlaps a proposed period of significant sedimentary cover in the region.

56

57

58 1.1. The Sääksjärvi impact structure, Finland

59 The Sääksjärvi impact structure is a deeply eroded, approximately 6 km in diameter feature

60 submerged beneath an eponymous lake in southwest Finland (Fig. 1; e.g., Willman et al., 2010).

61 Papunen (1969) suggested an impact origin for Sääksjärvi based on the observation of multiple

62 sets of closely spaced planar structures in quartz within heavily brecciated and vesiculated rocks

63 from the eastern shore of Lake Sääksjärvi and a glacial esker approximately 20 km to the

64 southeast. To the best of our knowledge, there are no known outcrops of in situ at

65 Sääksjärvi (e.g., Papunen 1969, 1973; Pihlaja and Kujala, 2000). The structure has been targeted

66 by six drill holes (totalling 860 m in length), including one into the centre of the gravity low

67 which penetrated 180 m of and other breccias (but not a coherent melt sheet) before

68 terminating in deformed mica gneiss basement at a depth below surface of 242 m (Pihlaja and

69 Kujala, 2000). Platinum group element patterns and Ni/Cr, Ni/Ir, and Cr/Ir ratios of impact melt

70 lithologies suggest the Sääksjärvi impactor was likely a non-magmatic iron (Schmidt et

71 al., 1997, Tagle et al. 2009).

72 Two studies have reported 40Ar/39Ar data for impact melt rocks and impact melt-bearing

73 breccias from Sääksjärvi. Bottomley et al. (1990) reported data from three samples described as

74 fine-grained melt rocks with quartz and feldspar clasts that were relatively unrecrystallised. The

75 samples displayed two distinct trends in their step-heating spectra: two samples gave

76 progressively older apparent ages from Ar release steps as heating continued (and 39Ar was

77 incrementally released), with ages rising from ca. 330 Ma and forming a plateau between 510

78 and 530 Ma, whereas aliquots from another sample gave a humped spectrum, with steps rising

79 from ca. 400 Ma to a peak at ca. 650 Ma and then forming a plateau at ca. 600 Ma. The dataset

80

81 82 Figure 1. Map of impact structures in the Baltic region. Best estimate ages, structure diameters and target rock

83 compositions from Schmieder and Kring (2020). When the age of an impact structure is given as a range in

84 Schmieder and Kring (2020) we use the midrange value – for example, for the Iso-Naakkima impact structure,

85 which is considered to have formed 1200–900 Ma, we use a value of 1050 Ma. Sääksjärvi is plotted using the new

86 608 ± 8 Ma age reported here. See Table 1 in Supplementary Appendix for full data used to create map.

87 Approximate extent of the Baltic Shield (including Caledonian Orogenic Belt) and Phanerozoic sediments after

88 Murrell and Andriessen (2004). The Caledonian foreland basin lay to the east of the Caledonian Orogenic Belt, over

89 present day Sweden and Finland, but its precise extent is uncertain (see text).

90

91 lacked any statistically relevant plateaus that may have represented an impact age (see discussion

92 of 40Ar/39Ar dating of impact structures by Jourdan [2012]) and Bottomley et al. (1990)

93 interpreted their data to indicate that a ≤330 Ma impact had incompletely degassed target rocks

94 with K-Ar ages of between 500 and 600 Ma. Müller et al. (1990) reported 40Ar/39Ar data for one

95 sample, which was described as a cryptocrystalline melt with >15 % clasts

96 (predominantly quartz and feldspar) and a matrix dominated by plagioclase and pyroxene. The

97 sample gave a humped age spectrum with apparent age domains of ca. 550 and ca. 600 Ma. With

98 no reason to prefer one age over the other, Müller et al. (1990) suggested that the ages should not

99 be taken “too seriously”. Nevertheless, Müller et al. (1990) considered a weighted average age of

100 560 ± 16 Ma to be their best estimate for the age of the . Similar to the dataset of

101 Bottomley et al. (1990), the data presented by Müller et al. (1990) lack statistically robust

102 plateaus and so none of the currently available 40Ar/39Ar data from impact melt at Sääksjärvi are

103 likely to record an accurate impact age.

104 1.2. Geological setting of the Sääksjärvi impact structure

105 The Sääksjärvi impact structure formed in crystalline basement of the Baltic Shield, in 1.9–

106 1.8 Ga rocks related to the (Fig. 2). During the Phanerozoic these

107 basement rocks were largely covered by sediments as a major foreland basin developed to the

108 east of the (Fig. 1) (e.g., Cederbom et al., 2000; Garfunkel and

109 Greiling, 1998; Larson et al., 1999; Middleton et al., 1996; Murrell and Andriessen, 2004).

110 However, the Caledonian foreland basin is not well preserved as rock exposures and estimates

111 for its extent and, particularly, life span vary significantly.

112 Fission-track thermochronology and inverse modelling suggests that most of Finland was

113 buried by approximately 1 km of Upper Paleozoic sediments that thickened to approximately 2.5

114 km towards the Caledonian border zone of western and southern Sweden (Cederbom et al., 2000;

115 Larson et al., 1999). Cederbom et al. (2000) presented models indicating that the basin existed

116 from the Devonian through to the early Mesozoic or even Cenozoic. Murrell and Andriessen

117 (2004) interpreted apatite fission-track data for southern Finland to indicate that late Silurian

118

119 Figure 2. Simplified geological map of the Sääksjärvi impact structure, Finland, after the 1:200 000 bedrock map of

120 the Geological Survey of Finland (GTK, 2017). White shaded area represents a lake (Sääksjärvi) and white lines

121 denote rivers.

122

123 heating, associated with the proposed Caledonian foreland basin, was followed by Cenozoic

124 cooling, representing exhumation and final exposure of the basement. Alternative models

125 suggest that even the deepest portions of the basin, in Sweden, were largely eroded by the mid-

126 Permian, with any Mesozoic sediment cover not exceeding 100 m (Söderlund et al., 2005).

127 1.3. Reconciling the Sääksjärvi impact event and the evolution of the Caledonian

128 foreland basin

129 The suggestion that Sääksjärvi, a small impact structure in crystalline basement with no

130 evidence for incorporation of supracrustal sedimentary rocks, formed at or after 330 Ma

131 (Bottomley et al., 1990) is problematic for models that imply the region was covered with

132 sediment for much of the Phanerozoic.

133 In light of the inconsistency between estimates for the timing of the Sääksjärvi impact event

134 based on 40Ar/39Ar data (Bottomley et al., 1990; Müller et al., 1990) and evidence that the

135 basement, which formed the target rock of the Sääksjärvi impact event, was in fact covered by up

136 to 1.5 km of sediments of the Caledonian foreland basin during a long period comprising much

137 of the Phanerozoic (e.g., Larson et al., 1999), we here revisit the age of the Sääksjärvi impact

138 event with high spatial resolution U–Pb geochronology. Impact ages have long been obtained

139 from U–Pb analysis of shocked accessory phases (e.g., Kamo and Krogh, 1995; Kamo et al.,

140 2011; Krogh et al., 1984, 1993, 1996; Mänttäri and Koivisto, 2001) with recent advances

141 offering the ability to better isolate impact-age domains of heterogeneous grains. These

142 developments include decreasing analytical volumes in in situ geochronology. Another critical

143 methodological advance is the ability to recognise impact-related microstructures that ultimately

144 can be dated (e.g., Cavosie et al., 2015; Erickson et al., 2016, 2017, 2020; Kenny et al., 2017,

145 2019; Timms et al., 2017). Promisingly, microstructures in zircon that are suitable for dating

146 (such as neoblasts in recrystallised grains) are not restricted to the largest impact structures but

147 have been documented at structures as small as the 1.2 km in diameter , USA

148 (Cavosie et al., 2016).

149 Here we report on imaging, microstructural analysis and U–Pb dating of shocked zircon from

150 the Sääksjärvi impact structure and present a new U–Pb age for the structure that is older than

151 previously suggested ages and consistent with the presence of the Caledonian foreland basin.

152 2. MATERIAL AND METHODS

153 Zircon grains were separated from five samples from the Sääksjärvi impact structure: three

154 samples of impact melt rock (n1047_IMP-2, n1049_IMP-1, and n5947_MS-17-1) and two

155 samples of impact melt-bearing breccia (n1048_IMP-3 and n1050_IMP-11b). Four of the

156 samples were taken from drill cores reported on by Pihlaja and Kujala (2000) and n5947_MS-17-

157 1 was a float sample. Unfortunately, the coordinates of the drill holes, as well as the intervals at

158 which these samples were collected, are not known and the precise origin of the float sample is

159 uncertain (although it was likely collected from the eastern or south-eastern shore of lake

160 Sääksjärvi or in glacial drift southeast of the structure – see Willman et al. [2010] and Fig. 1 in

161 Papunen [1969] for occurrences of impactites in the Sääksjärvi area).

162 The samples were disaggregated in a jaw crusher and ring-and-puck-style mill before their

163 heavy mineral fractions were concentrated by magnetic separation (with a hand magnet and

164 Frantz magnetic separator) and heavy liquid density separation (using methylene iodide with a

165 density of approximately 3.3 g/cm3). Zircon grains with a turbid, or cloudy, appearance were

166 preferentially picked from the heavy mineral separates as it has been shown that these are more

167 likely to be granular and contain age-reset domains than translucent grains (e.g., Schmieder et

168 al., 2015). The selected grains were placed on double-sided sticky tape and cast in 2.5 cm-

169 diameter epoxy mounts. The mounts were polished with a diamond suspension to expose grain

170 interiors and a final polish with colloidal silica prepared the grains for microstructural analysis.

171 This study comprised three separate analytical periods, in 2001 (period #1; the results of

172 which were presented by Mänttäri et al. [2004]), 2018 (period #2), and 2020 (period #3). In 2001

173 zircon grains from samples n1047_IMP-2, n1048_IMP-3, n1049_IMP-1, and n1050_IMP-11b

174 were imaged in scanning electron microscopy (SEM) backscattered electron (BSE) mode and

175 analysed for U–Pb isotopic composition and age by secondary ion mass spectrometry (SIMS) on

176 the Cameca IMS1270 ion microprobe at the NordSIMS Laboratory, Swedish Museum of Natural

177 History, with an analytical pit diameter of approximately 30 μm (generated with a duoplasmatron

178 O ion source). In 2018 zircon grains from sample n5947_MS-17-1 were imaged in BSE,

179 secondary electron (SE), and cathodoluminescence (CL) mode before undergoing U–Pb analysis

180 using the NordSIMS instrument upgraded to Cameca IMS1280, which employed an analytical

181 pit approximately 15 μm in diameter (also generated with a duoplasmatron O ion source). In

182 2020 grains from the 2001 analytical session were revisited. The grains were repolished to

183 remove the original analytical pits before being imaged in BSE, SE, and CL modes. Nine grains

184 with granular and/or porous textures were then selected for microstructural characterisation by

185 electron backscatter diffraction (EBSD) analysis (three of these had undergone U–Pb analysis in

186 2001 whereas the others had not been analysed previously). These grains then underwent U–Pb

187 isotopic analysis in which the analytical pit was approximately 5 μm in diameter (using a

188 Hyperion radio-frequency [RF] plasma ion source). Grain n1048_IMP-3_z5, which was analysed

189 with two 30 μm diameter pits in 2001 and nine 5 μm diameter pits in 2020, was subsequently

190 analysed with two 10 μm diameter analytical pits (also using the Hyperion RF plasma ion

191 source).

192 In analytical period #1, the polished surfaces of the zircon grains were imaged in BSE

193 mode. In periods #2-3, the grains were imaged in BSE, SE, and CL modes on an FEI Quanta

194 FEG 650 SEM equipped with a Gatan ChromaCL2 system at the Swedish Museum of Natural

195 History, Stockholm.

196 Electron backscatter diffraction analysis was performed with an Oxford Instruments

197 Nordlys detector attached to the FEI Quanta FEG 650 SEM. Grains were indexed for zircon,

198 reidite and monoclinic ZrO2 using match units based on crystallographic data from Hazen and

199 Finger (1979), Farnan et al. (2003), and Howard et al. (1988), respectively. Analytical conditions

200 and parameters include: accelerating voltage of 20 kV, working distance ~18 mm, stage tilt of

201 70°, electron backscatter pattern (EBSP) binning of 4x4, EBSP gain set to ‘High’, and

202 background defined with collection of 128 frames, Hough resolution set to 60, band detection

203 min/max of 6/8. Maps were collected with step sizes between 90 and 250 nm. Data collection

204 was performed in Oxford Instruments AZtec software and post-acquisition processing in Oxford

205 Instruments Channel 5 software v. 5.12. Data cleaning in Channel 5 comprised a wildspike

206 correction and a level six nearest neighbor zero solution extrapolation for all grains. Maps of

207 local misorientation represent the average misorientation between pixels in 5x5 (Fig. 3), 7x7

208 (Supplementary Figs 3-4), or 9x9 (Supplementary Fig. 5) grids. Grain boundaries, which are

209 shown in some maps, are here defined by >2° misorientation between adjacent pixels.

210 In the 2001 and 2018 U–Pb sessions, the analytical procedure largely followed the

211 routine of Whitehouse et al. (2001), with the duosplasmatron-generated primary beam yielding

212 analysis pits approximately 30 and 15 μm across, respectively. In the 2020 analytical period the

213 Hyperion-generated primary beam largely generated analysis pits approximately 5 μm across,

214 with two additional analyses performed with a 10 μm pit. A routine correction for common Pb

215 (Pbc) was applied during post-acquisition processing of the U–Pb data. The percentage of non-

206 206 216 radiogenic, i.e. common, Pb (f206%) in the total measured Pb was estimated from measured

204 217 Pb, assuming a modern-day Stacey and Kramers (1975) Pb composition, i.e. that Pbc

218 encountered in the analyses is likely to be modern and related to unavoidable sample

219 contamination (in crevices) during polishing of the grain mount (Zeck and Whitehouse, 1999). If

220 the Pbc is not modern but is actually ancient, U–Pb ages for high-Pbc analyses will be slightly

221 over-corrected, i.e. slightly too young. This is unlikely to be significant and Kenny et al. (2019)

222 reported that their concordia age for relatively Pbc-rich (f206% ranging from 0.15 to 3.08)

223 shocked zircon from the Lappajärvi impact structure would be less than 0.2 % older if the Pbc

224 encountered in those analyses was 1.9 Ga (a likely crystallisation age for those grains) and not

225 modern, as assumed. In all sessions 91500 standard zircon (ID-TIMS 206Pb/238U age = 1062.4 ±

226 0.8 Ma; Wiedenbeck et al., 1995) was used as the calibration reference material. Temora 2 (ID-

227 TIMS 206Pb/238U age = 416.78 ± 0.33 Ma; Black et al., 2004), which has consistently given ages

228 within ± 1% relative to the reference age when analysed at the NordSIMS laboratory (Jeon and

229 Whitehouse, 2015), was analysed as a secondary reference material during the 2018 session and

230 the 10 μm session in 2020, returning weighted average ages of 420.6 ± 3.0 Ma (MSWD = 0.93,

231 probability = 0.46, n = 6) and 420.6 ± 3.3 Ma (MSWD = 0.37, probability = 0.92, n = 8),

232 respectively. We used the decay constant values of Steiger and Jäger (1977) and all uncertainties

233 are presented at the 2σ level.

234 3. RESULTS

235 3.1. Imaging

236 Zircon grains from all five samples have diverse appearances with approximately half of the

237 deliberately picked cloudy grains exhibiting simple to complex internal zonation, with no

238 evidence for impact-related deformation, and half displaying granular and/or porous textures.

239 Of the apparently granular grains, some appear almost entirely recrystallised (e.g., Fig. 3) and

240 others retain conspicuous non-recrystallised domains (e.g., Fig. 4), whereas for others the degree

241 of the recrystallisation is not clear from imaging alone (e.g., Fig. 5). Individual neoblasts within

242 recrystallised grains range in size from less than 1 μm up to approximately 15 μm in dimension

243 and appear to take one of two forms: rounded granules 1-3 μm in size (e.g., Figs 3-4) or

244

245

246 Figure 3. Example of granular zircon from Sääksjärvi which preserves evidence for the former presence of reidite –

247 grain n1048_IMP-3_z13. F-G: Pole figures and plots of high-angle misorientation axes provide microstructural

248 evidence for the former presence of reidite in the form of: (i) domains of neoblasts systematically misoriented by

249 90°, (ii) coincidence among (001) and {110} poles, and (iii) high-angle misorientation axes coincident with poles to

250 {110}. Ellipses show locations of SIMS U–Pb analyses. IPF – Inverse Pole Figure.

251

252 Figure 4. Example of granular zircon from Sääksjärvi which does not preserve evidence for the former presence of

253 reidite – grain n1048_IMP-3_z4 (contrast with Fig. 3). Dashed ellipse A shows the approximate location of a 30 μm

254 analytical pit (session #1, 2001) prior to repolishing for imaging and microstructural analysis shown here.

255

256 Figure 5. A zircon grain (n1049_IMP-1_z11) from Sääksjärvi which appears granular in backscattered electron and

257 cathodoluminescence imaging but microstructural analysis reveals that apparent sub-domains share a single

258 approximate orientation.

259

260 subhedral to euhedral crystallites which are generally larger (e.g., Fig. 6). Some grains display

261 both small, rounded granules and larger, euhedral crystallites (e.g., Fig. 6).

262 Grains with pores generally also display some degree of recrystallisation elsewhere in the

263 grain (e.g., Fig. 6 and Supplementary Figs 1-2). Pores are usually round and approximately 0.2

264 μm in diameter; in some grains they are quite homogeneously distributed (e.g., Fig. 6) whereas in

265 others they occur in bands, suggestive of a possible relationship with pre-impact zoning within

266 the zircon (e.g., Supplementary Figs 1-2).

267 3.2. Microstructural characterisation

268 Six of the nine grains selected for EBSD analysis indexed well, producing whole-grain

269 microstructural maps (Figs. 3-5 and Supplementary Figs 3-5), whereas one grain only indexed on

270 approximately 20% of the exposed surface (Fig. 6) and two did not produce electron backscatter

271 diffraction patterns and so were not mapped (see imaging in Supplementary Figs 1-2).

272 In all grains zircon was the only phase to index. No evidence was encountered for ZrO2 (a

273 product of zircon dissociation at high temperature, which commonly accompanied granular

274 textures in impact-related zircon; e.g., Timms et al., 2017; Cavosie et al., 2018) or the current

275 presence of the high-pressure ZrSiO4 polymorph, reidite (e.g., Glass et al., 2002).

276 Microstructural characterisation reveals that many grains are almost entirely composed of

277 low-strain, newly crystallised domains, neoblasts (Fig. 3), whereas others retain non-

278 recrystallised domains defined by high degrees of strain apparent in maps of local misorientation

279 (e.g., Supplementary Figs 3,5). Of the seven grains which displayed at least some degree of

280 recrystallisation and were analysed by EBSD, four preserve crystallographic evidence for the

281 former presence of reidite in the form of: (i) domains of neoblasts systematically misoriented by

282

283 Figure 6. Shocked zircon that most accurately and precisely records the age of the Sääksjärvi impact event – grain

284 n1048_IMP-3_z5. The grain displays both granular (B) and porous (C) textures. Dashed ellipses A and B show the

285 approximate location of 30 μm analytical pits (session #1, 2001) prior to repolishing for imaging and microstructural

286 analysis shown here. Of the other analysis pits, C to K are approximately 5 μm in diameter and L and M are

287 approximately 10 μm in diameter. The 608 ± 8 Ma concordia age was obtained from data from analytical pits C, D,

288 G, and K, which are shown in red.

289 90°, (ii) coincidence among (001) and {110} poles, and (iii) high-angle misorientation axes

290 coincident with poles to {110} (Fig. 3, Supplementary Figs 3-5), which are indicative of the

291 transformation sequence zircon → reidite → zircon (Cavosie et al., 2016, 2018, Timms et al.,

292 2017). Of the three remaining granular grains, one displays a non-recrystallised core (with up to

293 5° internal misorientation) that is surrounded by granules with apparently random orientations

294 (Fig. 4), one appears granular in BSE and CL images but microstructural analysis shows that

295 sub-domains share a single approximate orientation (and deformation occurs in the form of

296 misorientation up to approximately 10° across the grain; Fig. 5), and the third is partly granular

297 and partly porous, with neoblasts not displaying any systematic relationships (Fig. 6).

298 3.3. U–Pb geochronology

299 The first U–Pb analyses, with analytical pit diameters of approximately 30 μm, were

300 performed in 2001 and target sites for the 43 analyses (on 34 zircon grains from four samples)

301 were based on BSE images. The data define two discordant trends in U–Pb space: one between

302 ca. 1.85 Ga and ca. 600 Ma and the other between ca. 600 Ma and present (Fig. 7A). The ca.

303 1.85 Ga ages are consistent with the Svecofennian ages of the target rocks at Sääksjärvi (Fig. 2)

304 whereas the ca. 600 Ma lower intercept age is not known from any rocks in the area. Most

305 analyses which plot to the right of the ca. 1.85 Ga to ca. 600 Ma discordant array in Tera-

306 Wasserburg (TW) space (Fig. 7A) have 207Pb/206Pb ages of ca. 600 Ma, consistent with later Pb

307 loss superimposed on grains which were fully age reset at ca. 600 Ma. However, a few grains

308 that lie off the ca. 1850–600 Ma trend have 207Pb/206Pb ages >600 Ma, indicating later Pb loss

309 superimposed on grains which were only partially reset at ca. 600 Ma.

310 The second batch of U–Pb analyses, with analytical pit diameters of approximately 15 μm,

311 was undertaken in 2018 and, for this analytical session, target sites for 22 analyses (on 17 zircon

312

313 Figure 7. Concordia diagrams displaying all secondary ion mass spectrometry (SIMS) U–Pb data gathered in this

314 study. A: Data generated with 30 and 15 μm analytical spots (using duoplasmatron ion source during SIMS

315 analyses). B: Data generated with 10 and 5 μm analytical spots (using Hyperion RF plasma ion source during SIMS

316 analyses). C: Data for grain n1048_IMP-3_z5 (see Fig. 5) – colours as in B. All data rendered at 2σ uncertainty.

317 grains from a previously unanalysed sample of impact melt rock, n5947_MS-17-1) were selected

318 based on BSE, SE, and CL images of the grains. Four analyses plot on or close to the concordia

319 curve at ca. 1850 Ma, consistent with the previously recognised Svecofennian ages; one analysis

320 lies on the ca. 1850–600 Ma discordia trend; and the remaining analyses plot to the right of the

321 discordia trend, again consistent with later Pb loss superimposed on the original discordant array.

322 Additionally, a single, almost concordant analysis gave a 207Pb/206Pb age of ca. 2.6 Ga (Fig. 7A),

323 highlighting an age not recognised in the earlier U–Pb dataset.

324 The third set of U–Pb analyses, most with analytical pit diameters of approximately 5 μm and

325 two with pit diameters of 10 μm, were performed in 2020. This time, target locations for 50

326 analyses (on nine zircon grains from three of the samples first analysed in 2001) were based on

327 EBSD maps and complementary BSE, SE, and CL images (except for the two grains which did

328 not index well and were targeted on the basis of traditional images alone; Supplementary Figs 1-

329 2). On the basis of microstructural characterisation, we targeted (i) individual neoblasts and

330 domains of neoblasts (with the aim to date recrystallisation), as well as (ii) non-recrystallised,

331 high-strain domains, and (iii) porous, non-indexing regions in three grains. Similar to earlier

332 analyses, the new dataset defines a trend between ca. 1.85 Ga and ca. 600 Ma with

333 approximately 15 analyses plotting to the right of the array in TW space, indicative of later Pb

334 loss superimposed on the original discordant array (Fig. 7B). However, many isotopic ratios and

335 ages from this session have significantly larger uncertainties than encountered previously – a

336 result of a higher proportion of analyses incorporating significant amounts of Pbc (40% of

337 analyses had f206% >1 compared to 14% of analyses in the 2001 session), likely due to targeting

338 of granular grains. One grain was particularly rich in Pbc, with calculated f206% values ranging

339 between 4.7 and 9.3 (grain n1050_IMP-11b_z18; Supplementary Fig. 4). The application of a

340 Pbc correction to the seven data points from this grain result in over-correction with six of the

341 Pbc-corrected data points lying below the concordia curve in TW space.

342 Plotting a regression through 56 of the 115 analyses from all analytical sessions (excluding

343 analyses which clearly lie off the ca 1.85 Ga to ca. 600 Ma trend, i.e. those with ages >2.0 Ga,

344 those which experienced significant later Pb loss, and those with substantial Pbc components)

345 yields a discordia line with a 1856 ± 15 Ma upper intercept and a 597 ± 7 Ma lower intercept

346 (mean square of weighted deviates [MSWD] = 2.1) (Supplementary Fig. 7). The relatively high

347 MSWD indicates some scatter in the data, which is likely related to subtle post-impact Pb loss

348 even in this screened population, and perhaps also heterogeneous pre-impact ages of the zircon

349 grains. Calculating a weighting mean 207Pb/206Pb age from the 21 analyses defining the

350 horizontal array between ca. 600 Ma and present day gives an age of 679 ± 18 Ma (MSWD =

351 3.4) (Supplementary Fig. 7). The relatively high MSWD on this age again indicates scatter in the

352 data, which, in this instance, can be attributed to variable degrees of Pb loss during the ca. 600

353 Ma event.

354 One grain in the 2020 dataset gave concordant data points at the ca. 600 Ma lower intercept

355 of the discordant array which lack significant Pbc contamination and therefore have small

356 uncertainties (Fig. 7B-C). Four concordant, low Pbc (f206% values of 0.01, 0.01, 0.02, and 0.06)

357 analyses on grain n1048_IMP-3_z5 give a concordia age (Ludwig, 1998) of 608 ± 8 Ma (2σ;

358 MSWD of concordance and equivalence = 1.06; probability of concordance and equivalence =

359 0.39) (Fig. 7C). Given the low Pbc values of these four analyses, whether the Pbc is modern, as

360 assumed in the Pbc correction, or ancient will not affect the age at this level of precision.

361 Significantly, the concordant data points with low Pbc were from the porous portion of this

362 mixed porous-granular grain (Fig. 6A,E and Supplementary Fig. 6; see also Supplementary Fig.

363 8 for a concordia diagram coloured according to zircon texture). Following the nine analyses

364 with 5 μm diameter analytical pits, this grain was targeted with two 10 μm diameter analytical

365 pits. However, neither of these two analyses were concordant, with one (targeting the porous

366 domain) plotting back along the discordia array and the other (targeting a large neoblast) having

367 experienced later Pb loss (Fig. 7B-C).

368 4. DISCUSSION

369 4.1. Age of the Sääksjärvi impact event

370 We interpret the ca. 1.85 Ga to ca. 600 Ma array of U–Pb data presented here to indicate

371 partial Pb loss during an Ediacaran impact into Palaeoproterozoic Svecofennian rocks of the

372 Baltic Shield. This was first suggested by Mänttäri et al. (2004) on the basis of the data acquired

373 in the 2001 analytical period. Within the context of the 115 analyses on 57 zircon grains from

374 five suevite and impact melt rock samples (over three analytical periods) the 608 ± 8 Ma

375 concordia age (Fig. 7C) from a porous zircon grain represents the best age estimate for the

376 Sääksjärvi impact event. We prefer the 608 ± 8 Ma age over the 597 ± 7 Ma lower intercept age

377 calculated from 56 analyses as the high MSWD of the latter indicates some scatter, which is

378 likely related to post-impact Pb loss in even that screened population. The 608 ± 8 Ma age is also

379 favourable to the weighted mean 207Pb/206Pb age calculated from the horizontal array as the older

380 age of the latter can be attributed to variable degrees of incomplete impact-induced Pb loss (as

381 reflected by its high MSWD value).

382 Our new age for the Sääksjärvi impact event is older than previous estimates of ≤330 Ma and

383 ca. 560 Ma, which were based on 40Ar/39Ar data, and makes Sääksjärvi the first confirmed

384 Ediacaran impact structure in Finland and the wider Baltic Shield (given that the 546 ± 5 Ma for

385 the Gardnos impact structure, Norway [Kalleson et al., 2009] is based on ambiguous U–Pb data

386 [e.g., Jourdan et al., 2009]). Furthermore, Sääksjärvi becomes only the second confirmed

387 Ediacaran impact structure worldwide, with the Acraman impact structure, Australia, constrained

388 to this time by the recognition of an associated ejecta layer in the Ediacaran Bunyeroo Formation

389 (Williams and Gostin, 2005; Schmieder et al., 2015, and references therein).

390 4.2. Implications for the Caledonian foreland basin

391 As originally proposed by Mänttäri et al. (2004), the new Ediacaran age for the Sääksjärvi

392 impact event reconciles the presence of this impact structure in crystalline basement with

393 regional sedimentary cover of the Caledonian foreland basin persisting from its Siluro-Devonian

394 onset through much of the Phanerozoic, and eliminates complications raised by a previously

395 proposed age of ≤330 Ma (Fig. 8). Furthermore, the ca. 330 Ma component in the 40Ar/39Ar data

396 of Bottomley et al. (1990) may have resulted from heating caused by burial related to Caledonian

397 sediment loading. The spread of 40Ar/39Ar data between ca. 330 Ma and ca. 650 Ma (Bottomley

398 et al., 1990) may now be reinterpreted as partial resetting of the 40Ar/39Ar system as a result of

399 burial and heating that is recognised in apatite fission-track datasets throughout the Baltic Shield

400 (e.g., Cederbom et al., 2000; Murrell and Andriessen, 2004; Larson et al., 1999). However, it is

401 not possible to ascertain whether the partial resetting related to burial and heating (caused by the

402 overburden of sediment) was superimposed on almost complete resetting of the 40Ar/39Ar system

403 at the 608 ± 8 Ma impact or if it simply represents partial resetting from predominantly

404 Palaeoproterozoic target rock ages. The interpretation that the 40Ar/39Ar data is recording

405 Paleozoic burial is also consistent with late Paleozoic lower-intercept ages of zircon across the

406 Baltic Shield, considered a result of Pb having been leached from the grains by hydrothermal

407 solutions related to the Caledonian Orogeny and foreland basin (Larson and Tullborg, 1998).

408 Similarly, a ca. 385 Ma age obtained from 40Ar/39Ar analysis of impact melt lithologies from the

409

410 Figure 8. Age distribution of impact structures in the Baltic region. A: Previous interpretations of 40Ar/39Ar data for

411 Sääksjärvi (shown in red) overlying the distribution of other impact structures formed in crystalline basement. B:

412 New U–Pb age from this study (shown in green) and distribution which includes the new age for Sääksjärvi. C: Age

413 data for all impact structures of the Baltic region considered in this study (see Fig. 1). Histogram coloured according

414 to scheme of Fig. 1. Five impact structures have poor age constraints and are not represented (see Table 1 in

415 Supplementary Appendix).

416

417 Gardnos impact structure, Norway, has been interpreted to represent thermal overprinting of the

418 Caledonian Orogeny and not the impact event itself (Grier et al., 1999), although this conclusion

419 has been contested (Jaret et al., 2016).

420 4.3. Implications for dating impact events with shocked zircon

421 It is striking that the four precise, concordant data points used to calculate the new age for the

422 Sääksjärvi impact event were from porous, non-indexing (i.e., poorly crystalline) domains of a

423 zircon grain and not from the many recrystallised, or granular, domains analysed.

424 A number of recent studies have, in addition to analysing recrystallised (neoblastic)

425 accessory phases such as zircon and apatite, also obtained impact ages from non-granular, but

426 pore-rich grains (e.g., McGregor et al., 2018, 2019; Schmieder et al., 2019; Schwarz et al., 2020).

427 However, most studies which have incorporated microstructural analysis have focused on

428 targeting recrystallised (neoblastic) domains of shocked minerals (e.g., Erickson et al., 2017,

429 2020; Kenny et al., 2017). An advantage of recrystallised grains is that, in addition to potentially

430 recording an impact age, they may also provide diagnostic evidence of shock if they record

431 crystallographic evidence for the former presence of reidite (Cavosie et al., 2016, 2018; Timms

432 et al., 2017). Porous textures, on the other hand, are not uniquely impact-related and, therefore,

433 an age determined from this texture in ex situ grains (for example, in a detrital sample) cannot be

434 considered a definitive impact age. We suggest that our findings at Sääksjärvi highlight a

435 potential pitfall of restricting analyses to a single textural type and that studies incorporating

436 microstructural analysis should not limit U–Pb analyses to recrystallised domains alone but

437 should also include porous textures if they are present.

438 Our data suggest that the use of an approximately 5 μm diameter analytical spot (enabled by

439 the recent installation of a Hyperion RF ion source at the NordSIMS laboratory) to isolate fully

440 age-reset domains within heterogeneous zircon grains is likely to significantly aid U–Pb dating

441 of impact events. Four of the nine 5 μm analytical spots on grain n1048_IMP-3_z5 gave

442 concordant, low-Pbc data points that enabled calculation of the 608 ± 8 Ma concordia age. In

443 contrast, none of the two 30 μm spots or two 10 μm spots on the same grain gave concordant

444 data. It therefore seems likely that if analyses were restricted to ≥ 10 μm, concordant data may

445 not have been obtained. We note that a number of recent studies have isolated impact-aged

446 domains with a 10 μm analytical spot (e.g., Erickson et al., 2017, 2020; Kenny et al., 2017, 2019;

447 Schwarz et al., 2020) but suggest that the advent of 5 μm spot analyses will aid targeting of

448 specific domains in complex grains, increase the likelihood of fully isolating impact-age regions,

449 and lead to more concordant data points (thereby improving uncertainties on impact ages).

450 Our new findings also reinforce the utility of U–Pb analysis of shocked accessory phases to

451 complement 40Ar/39Ar dating for small impact structures. Interpretation of 40Ar/39Ar spectra is

452 difficult for small impact structures, where the targets rock may not have been heated sufficiently

453 to completely degas radiogenic argon, 40Ar*, previously accumulated in the rock (e.g., Jourdan et

454 al., 2007; Pickersgill et al., 2020). Here, we show that complete resetting of the U–Pb system can

455 be achieved at structures as small as Sääksjärvi and note that is likely possible at even smaller

456 structures too, such as the 1.2 km-diameter Meteor Crater, Arizona, USA where recrystallised

457 zircon were documented (Cavosie et al., 2016). Additionally, 40Ar/39Ar dating is difficult at

458 structures for which there is a large time gap between target rock ages and the cratering event

459 (meaning that even minor degrees of excess 40Ar will significantly affect the calculated age;

460 Jourdan et al., 2007), but this is optimal for the U–Pb system in zircon whereby a large age

461 difference means that the impact-related discordant array intersects the concordia curve at a high

462 angle, allowing one to distinguish impact-related Pb loss from later, e.g., modern, Pb loss (e.g.,

463 Fig. 7A). Conversely, 40Ar/39Ar may be more applicable for craters for which there was a

464 relatively limited time gap between the target rock crystallisation age and the impact event (e.g.,

465 impacts into Phanerozoic target rocks), where U–Pb data will be distributed along the concordia

466 curve (as opposed to intersecting it), making it more complicated to reach unambiguous impact

467 ages (e.g., Erickson et al., 2017; Tohver et al., 2012). Finally, the higher closure temperature of

468 the U–Pb system in zircon relative to the 40Ar/39Ar system means that it may be more likely to

469 record the timing of the impact event and not later cooling of the impact structure (e.g., Kenny et

470 al., 2019; Schmieder and Kring, 2020) and that once the structure has cooled, zircon may be less

471 likely to be compromised by later heating, as appears to have affected the 40Ar/39Ar system at

472 Sääksjärvi.

473 4.4. Radiometric dating of small impact structures

474 Our new dataset highlights the potential of accurate and precise U–Pb dating of small impact

475 structures in crystalline basement to test, and even place boundaries on, models for the timing

476 and extent of palaeobasins that might otherwise be difficult to constrain. Although our new age

477 for the Sääksjärvi impact structure does not place tighter constraints on the timing of the

478 Caledonian foreland basin than were available from, for example, apatite fission-track studies, it

479 demonstrates the possibility that future radiometric dating of impact structures may do so. For

480 example, the Karikkoselkä impact structure, approximately 150 km northeast of Sääksjärvi, is a

481 1.5 km-diameter structure in crystalline basement that could provide new constraints, but its age

482 is still ambiguous: Pesonen et al. (1999) suggested that palaeomagnetic data are consistent with

483 ages of 230-260 Ma, 530-560 Ma, and 1650-1760 Ma, whereas Schmieder et al. (2010)

484 suggested a possible age as young as ca. 9 Ma on the basis of (U–Th)/He analyses of zircon.

485 More broadly, radiometric dating of small impact structures in crystalline basement

486 worldwide may provide constraints on sedimentary basins, the burial histories of

487 palaeolandscapes, and associated tectonics in less well-studied regions. In practise, this potential

488 utility will be hampered by the fact that most impact structures which formed on Earth have been

489 removed by erosion (e.g., Hergarten and Kenkmann, 2015; Johnson and Bowling, 2014), with

490 small impact structures more easily erased from the geological record than larger ones (which

491 produced a deeper record in the lithosphere). Today there are only approximately 110 known

492 impact structures on Earth that have diameters less than 6 km, the approximate size of Sääksjärvi

493 (Schmieder and Kring, 2020). Of these, approximately 60 are between 1.2 and 6 km in diameter,

494 meaning that at least this number are likely amenable to radiometric dating (given that datable

495 microstructures in zircon were identified at the 1.2 km in diameter Meteor Crater, USA [Cavosie

496 et al., 2016]). Encouragingly, most impact structures yet to be discovered on Earth’s surface are

497 likely to be less than 6 km in diameter (with as many as 90 in the 1 to 6 km range; Hergarten and

498 Kenkmann, 2015) and thus their radiometric ages may be of use in regional geological

499 reconstructions.

500 5. CONCLUSIONS

501 New U–Pb data for shocked zircon indicates that the Sääksjärvi impact structure, southwest

502 Finland, formed at 608 ± 8 Ma. This makes Sääksjärvi the first unambiguously dated Ediacaran

503 impact structure in the Baltic Shield, and one of only two confirmed Ediacaran impact structures

504 worldwide.

505 The 608 ± 8 Ma age for Sääksjärvi is older than previous estimates based on 40Ar/39Ar data

506 and confirms that the impact into exposed crystalline basement occurred prior to the

507 development of the Caledonian foreland basin. Apparently younger 40Ar/39Ar ages are likely a

508 result of Caledonian sediment loading and burial.

509 The accurate and precise dating of Sääksjärvi reinforces the utility of shocked accessory

510 phases in radiometric dating of small impact structures. Furthermore, the most accurate and

511 precise age for Sääksjärvi was acquired from porous, poorly crystalline zircon, indicating that

512 future studies should not focus on recrystallisation textures alone.

513 The use of a 5 μm diameter analytical spot (enabled by the Hyperion RF ion source) to

514 isolate fully age-reset domains within heterogeneous zircon grains is likely to significantly aid

515 U–Pb dating of impact events.

516 Dating of small impact structures (both those which are currently known and those yet to be

517 discovered) may help constrain the lifetime and spatial extent of sedimentary basins worldwide.

518 ACKNOWLEDGMENTS

519 We thank Kerstin Lindén for support with sample preparation , Heejin Jeon and Lev Ilyinsky for

520 assistance with ion microprobe analyses, Nick Timms and Timmons Erickson for constructive

521 reviews, and Deta Gasser for efficient editorial handling. This is NordSIMS Contribution No.

522 653 and Lunar and Planetary Institute Contribution No. 2386. The Lunar and Planetary Institute

523 is operated by Universities Space Research Association under a cooperative agreement with the

524 Science Mission Directorate of the National Aeronautics and Space Administration.

525

526 FUNDING

527 This project received funding from the European Union’s Horizon 2020 research and innovation

528 programme under the Marie Skłodowska-Curie Individual Fellowship Grant Agreement No.

529 792030, supporting GGK.

530

531 REFERENCES CITED

532 Black L. P., Kamo S. L., Allen C. M., Davis D. W., Aleinikoff J. N., Valley J. W., Mundil R.,

533 Campbell I. H., Korsch R. J., Williams I. S., and Foudoulis C. (2004) Improved

534 206Pb/238U microprobe geochronology by the monitoring of a trace-element-related matrix

535 effect; SHRIMP, ID–TIMS, ELA–ICP–MS and oxygen isotope documentation for a

536 series of zircon standards. Chemical 205, 115–140.

537 https://doi.org/10.1016/j.chemgeo.2004.01.003

538 Bottomley, R. J., York, D., and Grieve, R. A. F. (1990) 40Argon-39Argon dating of impact

539 craters. Proceedings of the Lunar and Planetary Science Conference 20, 421–431.

540 Buchner, E., Moilanen, J., Öhman, T., and Schmieder, M. (2009) Shock-Molten Sandstone

541 Clasts in Impact Melt Rocks: Age Constraints for the Paasselkä Impact Structure (SE

542 Finland). 40th Lunar and Planetary Science Conference, abstract no. 2169.

543 Cavosie A. J., Erickson T. M., Timms N. E., Reddy S. M., Talavera C., Montalvo S. D., Pincus

544 M. R., Gibbon R. J., and Moser D. (2015) A terrestrial perspective on using ex situ

545 shocked zircons to date lunar impacts. Geology 43, 999–1002.

546 https://doi.org/10.1130/g37059.1

547 Cavosie, A. J., Timms, N. E., Erickson, T. M., Hagerty, J. J., and Hörz, F. (2016)

548 Transformations to granular zircon revealed: Twinning, reidite, and ZrO2 in shocked

549 zircon from Meteor Crater (Arizona, USA). Geology 44, 703–706.

550 https://doi.org/10.1130/g38043.1

551 Cavosie, A. J., Timms, N. E., Ferrière, L., and Rochette, P. (2018) FRIGN zircon—The only

552 terrestrial mineral diagnostic of high-pressure and high-temperature shock deformation.

553 Geology 46, 891–894. https://doi.org/10.1130/G45079.1

554 Cederbom, C., Larson, S. Å., Tullborg, E.-L., and Stiberg, J.-P. (2000) Fission track

555 thermochronology applied to Phanerozoic thermotectonic events in central and southern

556 Sweden. Tectonophysics 316, 153–167. https://doi.org/10.1016/S0040-1951(99)00230-9

557 Erickson, T. M., Cavosie, A. J., Pearce, M. A., Timms, N. E., and Reddy, S. M. (2016) Empirical

558 constraints on shock features in monazite using shocked zircon inclusions: Geology 44,

559 635–638. https://doi.org/10.1130/G37979.1.

560 Erickson, T. M., Kirkland, C. L., Timms, N. E., Cavosie, A. J., and Davison, T. M. (2020)

561 Precise radiometric age establishes Yarrabubba, Western Australia, as Earth’s oldest

562 recognised meteorite impact structure. Nature Communications 11, 300.

563 https://doi.org/10.1038/s41467-019-13985-7

564 Erickson, T. M., Timms, N. E., Kirkland, C. L., Tohver, E., Cavosie, A. J., Pearce, M. A., and

565 Reddy, S.M. (2017) Shocked monazite chronometry: integrating microstructural and in

566 situ isotopic age data for determining precise impact ages. Contributions to Mineralogy

567 and Petrology 172, 11. https://doi.org/10.1007/s00410-017-1328-2

568 Farnan, I., Balan, E., Pickard, C. J., and Mauri, F. (2003) The effect of radiation damage on local

29 569 structure in the crystalline fraction of ZrSiO4: Investigating the Si NMR response to

570 pressure in zircon and reidite. American Mineralogist 88, 1663–1667.

571 https://doi.org/10.2138/am-2003-11-1205

572 Garfunkel, Z., and Greiling, R. O. (1998) A thin orogenic wedge upon thick foreland lithosphere

573 and the missing foreland basin. Geologische Rundschau 87, 314–325.

574 https://doi.org/10.1007/s005310050212

575 Glass, B. P., Liu, S., and Leavens, P. B. (2002) Reidite: An impact-produced high-pressure

576 polymorph of zircon found in marine sediments. American Mineralogist 87, 562–565.

577 https://doi.org/10.2138/am-2002-0420

578 Grier, J. A., Swindle, T. D., Kring, D. A., and Melosh, H. J. (1999) Argon-40/argon-39 analyses

579 of samples from the Gardnos impact structure, Norway. Meteoritics and Planetary

580 Science 34, 803–807. https://doi.org/10.1111/j.1945-5100.1999.tb01393.x

581 Grieve, R. A. F. (2006) Impact structures in Canada. Geological Association of Canada GEOtext

582 Number 5, 210 pp. ISBN: 978-1-897095-11-9

583 Grieve, R. A. F., and Head III, J. W. (1983) The Manicouagan impact structure: An analysis of

584 its original dimensions and form. Journal of Geophysical Research: Solid Earth 88,

585 A807–A818. https://doi.org/10.1029/JB088iS02p0A807

586 GTK (2017) Bedrock of Finland 1:200 000, v. 2.1 https://gtkdata.gtk.fi/Maankamara/index.html

587 accessed: 16th January 2020.

588 Hazen, R. M., and Finger, L. W. (1979) Crystal structure and compressibility of zircon at high

589 pressure. American Mineralogist 64, 196–201.

590 Hergarten, S., and Kenkmann, T. (2015) The number of impact craters on Earth: Any room for

591 further discoveries? Earth and Planetary Science Letters 425, 187–192.

592 http://dx.doi.org/10.1016/j.epsl.2015.06.009

593 Howard, C. J., Hill, R. J., and Reichert, B. E. (1988) Structures of ZrO2 polymorphs at room

594 temperature by high‐resolution neutron powder diffraction. Acta Crystallographica

595 Section B 44, 116–120. https://doi.org/10.1107/S0108768187010279

596 Jaret, S. J., Hemming, S. R., Rasbury, E. T., Ramezani, J., and Jacobsen, S. B. (2016) The case

597 for an impact at the Gardnos impact structure at 385 Ma. 79th Annual Meeting of the

598 Meteoritical Society, abstract no. 6476.

599 Jeon, H., and Whitehouse, M. J. (2015) A critical evaluation of U–Pb calibration schemes used

600 in SIMS zircon geochronology. Geostandards and Geoanalytical Research 39, 443–452.

601 https://doi.org/10.1111/j.1751-908X.2014.00325.x

602 Johnson, B. C., and Bowling, T. J. (2014) Where have all the craters gone? Earth’s bombardment

603 history and the expected terrestrial cratering record. Geology 42, 587–590.

604 https://doi.org/10.1130/G35754.1

605 Jourdan, F. (2012) The 40Ar/39Ar dating technique applied to planetary sciences and terrestrial

606 impacts. Australian Journal of Earth Sciences 59, 199–224.

607 https://doi.org/10.1080/08120099.2012.644404

608 Jourdan, F., Renne, P. R., and Reimold, W. U. (2007) The problem of inherited 40Ar* in dating

609 impact glass by the 40Ar/39Ar method: Evidence from the Tswaing (South

610 Africa). Geochimica et Cosmochimica Acta 71, 1214–1231.

611 https://doi.org/10.1016/j.gca.2006.11.013

612 Jourdan, F., Renne, P. R., and Reimold, W. U. (2009) An appraisal of the ages of terrestrial

613 impact structures. Earth and Planetary Science Letters 286, 1–13.

614 https://doi.org/10.1016/j.epsl.2009.07.009

615 Kalleson, E., Corfu, F., and Dypvik, H. (2009) U–Pb systematics of zircon and titanite from the

616 Gardnos impact structure, Norway: Evidence for impact at 546 Ma? Geochimica et

617 Cosmochimica Acta 73, 3077–3092. https://doi.org/10.1016/j.gca.2009.02.020

618 Kamo, S. L., and Krogh, T. E. (1995) source for shocked zircon crystals from

619 the Cretaceous-Tertiary boundary layer, Saskatchewan: Evidence from new U-Pb data.

620 Geology 23, 281–284. https://doi.org/10.1130/0091-

621 7613(1995)023<0281:CCSFSZ>2.3.CO;2

622 Kamo, S. L., Lana, C., and Morgan, J. V. (2011) U–Pb ages of shocked zircon grains link distal

623 K–Pg boundary sites in Spain and Italy with the Chicxulub impact. Earth and Planetary

624 Science Letters 310, 401–408. https://doi.org/10.1016/j.epsl.2011.08.031

625 Kenny, G. G., Morales, L. F., Whitehouse, M. J., Petrus, J. A., and Kamber, B. S. (2017) The

626 formation of large neoblasts in shocked zircon and their utility in dating impacts.

627 Geology 45, 1003–1006. https://doi.org/10.1130/G39328.1

628 Kenny, G. G., Schmieder, M., Whitehouse, M. J., Nemchin, A. A., Morales, L. F., Buchner, E.,

629 Bellucci, J. J., and Snape, J. F. (2019) A new U-Pb age for shock-recrystallised zircon

630 from the Lappajärvi impact crater, Finland, and implications for the accurate dating of

631 impact events. Geochimica et Cosmochimica Acta 245, 479–494.

632 https://doi.org/10.1016/j.gca.2018.11.021

633 Krogh, T. E., Davis, D. W., and Corfu, F. (1984) Precise U-Pb zircon and baddeleyite ages for

634 the Sudbury area. In: Pye, E. G., Naldrett, A. J., and Giblin, P. E. (eds) The Geology and

635 Ore Deposits of the Sudbury Structure. Ontario Geological Survey, Special Volume 1,

636 431–446.

637 Krogh T. E., Kamo S. L., and Bohor B. F. (1993) Fingerprinting the K/T impact site and

638 determining the time of impact by U-Pb dating of single shocked zircons from distal

639 ejecta. Earth and Planetary Science Letters 119, 425–429. https://doi.org/10.1016/0012-

640 821X(93)90150-8

641 Krogh T. E., Kamo S. L., and Bohor B. F. (1996) Shock metamorphosed zircons with correlated

642 U-Pb discordance and melt rocks with concordant protolith ages indicate an impact origin

643 for the Sudbury structure. In: Basu, A., and Hart, S. (eds) Earth processes: reading the

644 isotopic code. American Geophysical Union 95, 343–353.

645 arson, S. ., and Tullborg, E.-L. (1998) Why Baltic Shield zircons yield late Paleozoic, lower-

646 intercept ages on U-Pb concordia. Geology 26, 919–922. https://doi.org/10.1130/0091-

647 7613(1998)026<0919:WBSZYL>2.3.CO;2

648 Larson, S. Å., Tullborg, E.-L., Cederbom, C., and Stiberg, J.-P. (1999) Sveconorwegian and

649 Caledonian foreland basins in the Baltic Shield revealed by fission-track

650 thermochronology. Terra Nova 11, 210–215. https://doi.org/10.1046/j.1365-

651 3121.1999.00249.x

652 Lehnert, O., Meinhold, G., Bergström, S. M., Calner, M., Ebbestad, J. O. R., Egenhoff, S., Frisk,

653 Å. M., Hannah, J. L., Högström, A. E. S., Huff, W. D., Juhlin, C., Maletz, J., Stein, H. J.,

654 Sturkell, E., and Vandenbroucke, T. R. A. (2012) New –Silurian drill cores

655 from the Siljan impact structure in central Sweden: an integral part of the Swedish Deep

656 Drilling Program. GFF 134, 87–98. https://doi.org/10.1080/11035897.2012.692707

657 ehnert, O., Meinhold, G., Wu, R., Calner, M., and Joachimski, M. M. (2014) δ13C

658 chemostratigraphy in the upper Tremadocian through lower Katian (Ordovician)

659 carbonate succession of the Siljan district, central Sweden. Estonian Journal of Earth

660 Sciences 63, 277–286. https://doi.org/10.3176/earth.2014.31

661 Ludwig, K. R. (1998) On the treatment of concordant uranium-lead ages. Geochimica et

662 Cosmochimica Acta 62, 665–676. https://doi.org/10.1016/S0016-7037(98)00059-3

663 Mänttäri, I., Kohonen, J., Kujala, H., and Pihlaja, P. (2004) A revised age for the Sääksjärvi

664 meteorite impact, southwestern Finland: The connexion with a Caledonian foreland

665 basin. 32nd International Geological Congress (Florence, Italy), abstract no. 1434.

666 Mänttäri, I., and Koivisto, M. (2001) Ion microprobe uranium-lead dating of zircons from the

667 Lappajärvi impact crater, western Finland. Meteoritics and Planetary Science 36, 1087–

668 1095. https://doi.org/10.1111/j.1945-5100.2001.tb01946.x

669 McGregor, M., McFarlane, C. R. M., and Spray, J. G. (2018) In situ LA-ICP-MS apatite and

670 zircon U–Pb geochronology of the Nicholson Lake impact structure, Canada: Shock and

671 related thermal effects. Earth and Planetary Science Letters 504, 185–197.

672 https://doi.org/10.1016/j.epsl.2018.10.006

673 McGregor, M., McFarlane, C. R. M., and Spray, J. G. (2019) In situ multiphase U–Pb

674 geochronology and shock analysis of apatite, titanite and zircon from the Lac La

675 Moinerie impact structure, Canada. Contribution to Mineralogy and Petrology 174, 62.

676 https://doi.org/10.1007/s00410-019-1598-y

677 Middleton, M. F., Tullborg, E.-L., Larson, S. Å., and Björklund, L. (1996) Modelling of a

678 Caledonian foreland basin in Sweden: petrophysical constraints. Marine and Petroleum

679 Geology 13, 407–413. https://doi.org/10.1016/0264-8172(95)00087-9

680 Müller, N., Hartung, J. B., Jessberger, E. K., and Reimold, W. U. (1990) 40Ar‐39Ar ages of

681 , Jänisjärvi, and Sääksjärvi impact craters. Meteoritics 25, 1–10.

682 https://doi.org/10.1111/j.1945-5100.1990.tb00964.x

683 Murrell, G. R., and Andriessen, P. A. M. (2004) Unravelling a long-term multi-event thermal

684 record in the cratonic interior of southern Finland through apatite fission track

685 thermochronology Physics and Chemistry of the Earth 29, 695–706.

686 https://doi.org/10.1016/j.pce.2004.03.007

687 Papunen, H. (1969) Possible impact metamorphic textures in the erratics of the Lake Sääksjärvi

688 area in southern Finland. Bulletin of the Geological Society of Finland 41, 151-155.

689 Papunen, H. (1973) Chemical composition and origin of the shock metamorphic rocks of the

690 Sääksjärvi area, Finland. Bulletin of the Geological Society of Finland 45, 29–34.

691 Pesonen, L. J., Elo, S., Lehtinen, M., Jokinen, T., Puranen, R., and Kivekäs, L. (1999) Lake

692 Karikkoselkä impact structure, central Finland: new geophysical and petrographic results.

693 In: Dressler, B. O., and Sharpton, V. L. (eds) Large Meteorite Impacts and Planetary

694 Evolution II, GSA Special Paper 339, Geological Society of America, Boulder, CO, pp

695 131–148.

696 Pickersgill, A. E., Mark, D. F., Lee, M. R., and Osinski, G. R. (2020) 40Ar/39Ar systematics of

697 melt lithologies and target rocks from the Gow Lake impact structure, Canada.

698 Geochimica et Cosmochimica Acta 274, 317–332.

699 https://doi.org/10.1016/j.gca.2020.01.025

700 Pihlaja, P. J., and Kujala, H. (2000) Some geochemical features of impact melts, and

701 target rocks at Sääksjärvi impact structure, Finland. In: Plado, J., and Pesonen, L. J. (eds)

702 Programme and Abstracts of ESF workshop: Meteorite Impacts in Precambrian Shields.

703 Geological Survey of Finland and University of Helsinki, Espoo: p 50

704 Puura, V., and Plado, J. (2005). Settings of meteorite impact structures in the Svecofennian

705 crustal domain. In: Koeberl, C., and Henkel, H. (eds) Impact tectonics. Springer, Berlin,

706 Heidelberg.: pp. 211–245.

707 Schmidt, G., Palme, H., and Kratz, K. L. (1997) Highly siderophile elements (Re, Os, Ir, Ru, Rh,

708 Pd, Au) in impact melts from three European impact craters (Sääksjärvi, , and

709 Dellen): Clues to the nature of the impacting bodies. Geochimica et Cosmochimica Acta,

710 61, 2977–2987. https://doi.org/10.1016/S0016-7037(97)00129-4

711 Schmieder, M., and Kring, D. A. (2020) Earth's Impact Events Through Geologic Time: A List

712 of Recommended Ages for Terrestrial Impact Structures and Deposits. Astrobiology 20,

713 91–141. https://doi.org/10.1089/ast.2019.2085

714 Schmieder, M., Buchner, E., Jourdan, F., Schwarz, W. H., Trieloff, M., van Soest, M. C.,

715 Wartho, J. A., Hodges, K. V., Moilanen, J., Hietala, S., and Öhman, T. (2010) Updating

716 the Finnish Impact Cratering Record. 41st Lunar and Planetary Science Conference,

717 abstract no. 2036.

718 Schmieder, M., Tohver, E., Jourdan, F., Denyszyn, S. W., and Haines, P. W. (2015) Zircons

719 from the Acraman impact melt rock (South Australia): , U–Pb and

720 40Ar/39Ar systematics, and implications for the isotopic dating of impact events.

721 Geochimica et Cosmochimica Acta 161, 71–100.

722 https://doi.org/10.1016/j.gca.2015.04.021

723 Schmieder, M., Shaulis, B. J., Lapen, T. J., Buchner, E., and Kring, D. A. (2019) In situ U–Pb

724 analysis of shocked zircon from the Charlevoix impact structure, Québec, Canada.

725 Meteoritics and Planetary Science 54, 1808–1827.

726 https://doi.org/10.1111/maps.13315

727 Schwarz, W. H., Hanel, M., and Trieloff, M. (2020) U‐Pb dating of zircons from an impact melt

728 of the Nördlinger Ries crater. Meteoritics and Planetary Science 13437.

729 https://doi.org/10.1111/maps.13437

730 Slater, B. J., and Willman, S., (2019) Early Cambrian small carbonaceous (SCFs) from an

731 impact crater in western Finland. Lethaia 52, 570–582. https://doi.org/10.1111/let.12331

732 Stacey, J. S., and Kramers, J. D. (1975) Approximation of terrestrial lead isotope evolution by a

733 two-stage model. Earth and Planetary Science Letters 26, 207–221.

734 https://doi.org/10.1016/0012-821X(75)90088-6

735 Steiger, R. H., and Jäger, E. (1977) Subcommission on geochronology: convention on the use of

736 decay constants in geo-and cosmochronology. Earth and Planetary Science Letters 36,

737 359–362. https://doi.org/10.1016/0012-821X(77)90060-7

738 Söderlund, P., Juez-Larré, J., Page, L. M., and Dunai, T. J. (2005) Extending the time range of

739 apatite (U–Th)/He thermochronometry in slowly cooled terranes: Palaeozoic to Cenozoic

740 exhumation history of southeast Sweden. Earth and Planetary Science Letters 239, 266–

741 275. https://doi.org/10.1016/j.epsl.2005.09.009

742 Tagle, R., Schmitt, R. T., and Erzinger, J. (2009) Identification of the projectile component in the

743 impact structures Rochechouart, France and Sääksjärvi, Finland: implications for the

744 impactor population for the earth. Geochimica et Cosmochimica Acta 73, 4891–4906.

745 https://doi.org/10.1016/j.gca.2009.05.044

746 Timms, N. E., Erickson, T. M., Pearce, M. A., Cavosie, A. J., Schmieder, M., Tohver, E., Reddy,

747 S. M., Zanetti, M. R., Nemchin, A. A., and Wittmann, A. (2017) A pressure-temperature

748 phase diagram for zircon at extreme conditions. Earth-Science Reviews 165, 185–202.

749 https://doi.org/10.1016/j.earscirev.2016.12.008

750 Tohver E., Lana C., Cawood P., Fletcher I., Jourdan F., Sherlock S., Rasmussen B., Trindade R.,

751 Yokoyama E., Souza Filho C., and Marangoni, Y. (2012) Geochronological constraints

752 on the age of a Permo–Triassic impact event: U–Pb and 40Ar/39Ar results for the 40 km

753 Araguainha structure of central Brazil. Geochimica et Cosmochimica Acta 86, 214–227.

754 https://doi.org/10.1016/10.1016/j.gca.2012.03.005

755 Wiedenbeck M., Allé P., Corfu F., Griffin W. L., Meier M., Oberli F., Quadt A. V., Roddick J.

756 C., and Spiegel W. (1995) Three natural zircon standards for U-Th-Pb, Lu-Hf, trace

757 element and REE analyses. Geostandards Newsletter 19, 1–23.

758 https://doi.org/10.1111/j.1751-908X.1995.tb00147.x

759 Williams, G. E., and Gostin, V. A. (2005) Acraman – Bunyeroo impact event (Ediacaran), South

760 Australia, and environmental consequences: twenty-five years on. Australian Journal of

761 Earth Sciences 52, 607–620. https://doi.org/10.1080/08120090500181036

762 Willman, S., Plado, J., Raukas, A., and Bauert, H. (2010) Meteorite Impact Structures –

763 Geotourism in the Central Baltic. NGO Geoguide Baltoscandia, Tallinn, Estonia, 80 pp.

764 ISBN: 978-9985-9973-2-1

765 Zeck, H. P., and Whitehouse, M. J. (1999) Hercynian, Pan-African, Proterozoic and Archean

766 ion-microprobe zircon ages for a Betic-Rif core complex, Alpine belt, W Mediterranean –

767 consequences for its P-T-t path. Contributions to Mineralogy and Petrology 134, 134–

768 149. https://doi.org/10.1007/s004100050474

769

770 Supplementary data

771  Supplementary Table 1 (Data used to produce map in Fig. 1 in main text)

772  Supplementary Figure 1 (Imaging of grain n1049_IMP-1_z4)

773  Supplementary Figure 2 (Imaging of grain n1049_IMP-1_z7)

774  Supplementary Figure 3 (Imaging and EBSD analysis of grain n1050_IMP-11b_z16)

775  Supplementary Figure 4 (Imaging and EBSD analysis of grain n1050_IMP-11b_z18)

776  Supplementary Figure 5 (Imaging and EBSD analysis of grain n1050_IMP-11b_z14)

777  Supplementary Figure 6 (Post-analysis imaging of grain n1048_IMP-3_z5)

778  Supplementary Figure 7 (Concordia diagram with regression through sub-population)

779  Supplementary Figure 8 (Concordia diagram coloured according to zircon texture)

780

781

Name Longitude Latitude Country Size (km) Age (Ma) Target rock Notes

Dellen 16.70 61.86 Sweden 19 141 Crystalline Gardnos 09.00 60.65 Norway 5 546 Crystalline The U–Pb data for Gardnos are considered equivocal. Granby 14.93 58.42 Sweden 3 466 Sedimentary Hummeln 16.25 57.37 Sweden 1.2 465 Mixed Iso-Naakkima 27.15 62.18 Finland 3 1050 Sedimentary Age range: 1200-900 Ma Jänisjärvi 30.92 61.97 Russia 14 687 Crystalline Kaali 22.67 58.40 Estonia 0.11 0.00324 Sedimentary Impact into unconsolidated material Kärdla 22.77 59.02 Estonia 4 455 Mixed Karikkoselkä 25.25 62.22 Finland 1.5 Crystalline Age range 260-230 Ma, or 9 Ma Keurusselkä 24.60 62.13 Finland 30 1151 Crystalline Lappajärvi 23.67 63.15 Finland 23 78 Mixed Lockne 14.82 63.00 Sweden 10.75 455 Mixed Diameter range: 7.5-14km 20.13 60.15 Finland 9 Mixed Age < 458 Ma Målingen 14.55 62.92 Sweden 0.7 455 Mixed Mien 14.86 56.42 Sweden 9 122 Crystalline Neugrund 23.67 59.33 Estonia 20 535 Crystalline Age range: 540-530 Ma Paasselkä 29.08 62.03 Finland 10 231 Mixed Ritland 06.44 59.23 Norway 2.7 520 Mixed Age range: 540-500 Ma Sääksjärvi 22.40 61.40 Finland 6 608 Crystalline Saarijärvi 28.38 65.28 Finland 1.5 560 Crystalline Age range: 600-520 Ma Siljan 14.87 61.03 Sweden 52 381 Mixed Söderfjärden 21.58 63.03 Finland 6.6 Crystalline Age range: 1880-640 Ma Summanen 25.38 62.65 Finland 2.6 Crystalline Age < 1880 Ma Suvasvesi N 28.17 62.70 Finland 3.5 85 Crystalline S 28.22 62.60 Finland 3.8 Crystalline Age range: 1880-710 Ma Tvären 17.42 58.77 Sweden 2 458 Mixed

Within bounds of map but not considered

Dobele 23.25 56.58 Latvia Ilumetsä 27.42 57.97 Estonia Mishina Gora 28.05 58.72 Russia Vepriai 24.58 55.08 Lithuania 782

783 Supplementary Table 1. Data used to produce map in Fig. 1 in main text. Data from Schmieder and Kring (2020),

784 except age for the Sääksjärvi impact structure which is from this study.

785 786 Supplementary Figure 1. Imaging of grain n1049_IMP-1_z4.

787 788 Supplementary Figure 2. Imaging of grain n1049_IMP-1_z7.

789 790

791 792 Supplementary Figure 3. Imaging and EBSD analysis of grain n1050_IMP-11b_z16. 793

794 795 Supplementary Figure 4. Imaging and EBSD analysis of grain n1050_IMP-11b_z18.

796 797 Supplementary Figure 5. Imaging and EBSD analysis of grain n1050_IMP-11b_z14. 798

799 800 Supplementary Figure 6. Post-analysis backscattered electron imaging of grain n1048_IMP-3_z5. The bright

801 material surrounding the grain and in some analytical pits and vesicles is gold remnant from the coating for ion

802 microprobe analysis. Pits would have been free from gold after analysis but some was trapped in a number of the

803 pits during brief polishing to remove the coating from the rest of the grain. No evidence for the presence of ZrO2

804 was encountered. The 608 ± 8 Ma concordia age was calculated from data from analytical pits C, D, G, and K.

805 806 Supplementary Figure 7. Tera-Wasserburg concordia diagram with a regression through the sub-population

807 defining the ca. 1850 to ca. 600 Ma trend and a weighted mean 207Pb/206Pb age of analyses which record later Pb

808 loss. MSWD – mean square of weighted deviates. 809

810 811 Supplementary Figure 8. Tera-Wasserburg concordia diagram coloured according to zircon texture. Note that

812 information on texture is not available for analytical period #1 (undertaken in 2001) and so these data are not

813 presented.

814