Tectonophysics 580 (2012) 88–99 Contents lists available at SciVerse ScienceDirect Tectonophysics j ourna l homepage: www.e lsev ie r .com/ locate / tectoA new interpretation of the sedimentary cover in the western Siljan Ring area, central Sweden, based on seismic data Christopher Juhlin a,⁎, Erik Sturkell b, Jan Ove R. Ebbestad c, Oliver Lehnert d, Anette E.S. Högström e, Guido Meinhold f a Department of Earth Sciences–Geophysics, Uppsala University, Villavägen 16, SE‐752 36 Uppsala, Sweden b Department of Earth Sciences, University of Gothenburg, PO Box 460, SE-405 30 Göteborg, Sweden c Museum of Evolution, Uppsala University, Norbyvägen 16, SE‐752 36 Uppsala, Sweden d GeoZentrum Nordbayern, Universität Erlangen, Schloßgarten 5, D-91054 Erlangen, Germany e Tromsø University Museum, Natural Sciences, N-9037 Tromsø, Norway f Sedimentology and Environmental Geology, Geoscience Centre, University of Göttingen, Goldschmidtstrasse 3, D-37077 Göttingen, Germany⁎ Corresponding author. Tel.: +46 18 4712392; fax: + E-mail address: Christopher.Juhlin@geo.uu.se (C. Juh 0040-1951 © 2012 Elsevier B.V. http://dx.doi.org/10.1016/j.tecto.2012.08.040 Open access under CC BYa b s t r a c ta r t i c l e i n f oArticle history: Received 22 May 2012 Received in revised form 14 August 2012 Accepted 27 August 2012 Available online 1 September 2012 Keywords: Impact crater Silurian Ordovician Reflection seismic Refraction seismic BoreholeTwo new reflection seismic profiles over the Paleozoic successions of the western part of the Siljan Ring impact structure show a contrasting seismic signature. The more southerly c. 10 km long Mora profile reveals a highly disturbed structure, with only a few kilometers of relatively horizontally layered structures observed. However, interpretations of refracted arrivals in the data, that can be correlated to reflections, indicate the Silurian clastic rocks to be about 200 m thick in the central part of the profile.Weak reflections from about 600 mdepth suggest a 400 m thick Ordovician limestone sequence to be present. Cores from the area show a mainly shale lithology for the Silurian and only a thin sequence of Ordovician strata, suggesting a rapid thickening of the Ordovician towards the north. On the more northern c. 12 km Orsa profile clear reflections from the Paleozoic successions are seen along the entire profile, except on the southernmost few kilometers. Based on interpretations of refracted arrivals, the Silurian succession appears to be considerably thinner here, and possibly absent at some locations. The Ordovician is also interpreted to be thinner in this area, with a maximum thickness of about 200–300 m along most of the profile. A deeper reflection from about 2 km within the crystalline basement may represent a dolerite sill. The lack of clear basement reflections on the Mora profile can be attributed to near-surface conditions and the acquisition geometry. The seismic data and recent coring in the area suggest the presence of a deeper paleo-basin towards the southwest with significantly more shales being deposited and the Paleozoic successions being severely disturbed. The shallow coring and seismic data will help form the basis for locating future boreholes for deeper drilling to study impact processes and the Paleozoic evolution of central Sweden. © 2012 Elsevier B.V. Open access under CC BY-NC-ND license.1. Introduction During the summer of 2011 two reflection seismic profiles were acquired in the western part of the Siljan impact structure (Fig. 1). The work was conducted under the framework of the project ‘Concentric Impact Structures in the Paleozoic’ (CISP), which aims to study impact mechanics, tectonics and the paleobiology of the Siljan and Lockne impact structures. CISP is financed 2010–2012 by the Swedish Research Council (VR) as one of two currently VR funded projects that are part of the Swedish Deep Drilling Program (SDDP; Lorenz, 2010). To date, the overall objective for the CISP studies has been to establish reliable geo- logical models for the planned deeper drilling in the Siljan and Lockne46 18 501 110. lin). -NC-ND license.impact structures under the SDDP umbrella; for an outline of detailed scientific objectives see Högström et al. (2010). Acquisition of new reflection seismic data has been the first step in defining the drilling locations in the Siljan impact crater. Previously acquired seismic data in the area were obtained in conjunction with the search for abiogenic methane in the 1980s and early 1990s (Bodén and Erikson, 1988). A total of nine lines were acquired, mostly in the eastern and northern parts of the Siljan Ring area, but one of them, Line 9, extended from the central part of the crater nearly to the town of Mora (Fig. 1). These profiles show the upper crustal Transcandinavian Igneous Belt (TIB) granitic rock in the area to be fairly transparent, but with strong high amplitude sub-horizontal reflections at distinct levels in the upper 10 km. Deep drilling within the crater rim revealed these high amplitude reflections to be generated by relatively thin (5–50 m) dolerite sills (Juhlin, 1990; Papasikas and Juhlin, 1997). A deeper reflec- tive zone is present at about 12–15 km throughout the area (Juhojuntti Precambrian bedrock 0 10 km TownsLate Silurian–Devonian sediments (Orsa Sandstone) Ordovician sediments Igrene drill sitesKullsberg Mounds Silurian sediments Legend Transpression ridges Siljan SWEDEN 600km N Solberga 1 Boda Mounds Stumsnäs 1 Mora 1 N Lake Ore Lake Skattungen Mora Orsa Lake Orsa Rättvik Lake Siljan Key sections Amtjärn Fjäcka Kårgärde Or sa pro file Mora profile Seismic profiles Styggforsen Skattungbyn Fig. 1. Location map of the Siljan Ring, showing distribution of sedimentary rocks, key geological sections, the drill sites for cores taken by Igrene AB, the position of the new seismic profiles, and the transpression ridges of Kenkmann and von Dalwigk (2000a,b). Modified from Ebbestad and Högström (2007). 89C. Juhlin et al. / Tectonophysics 580 (2012) 88–99and Juhlin, 1998) thatmayhave adifferent origin. North of the SiljanRing, where Svecofennian Domain rocks are present at the surface, relatively strong lower crustal reflectivity is observed with the Moho interpreted to be at about 45 km (Juhojuntti and Juhlin, 1998). In this paper we present results from two new seismic profiles that mainly pass over the Paleozoic sedimentary successions of the western part of the Siljan Ring. One profile, about 10 km long and striking in the W–E direction, was located just north of the town of Mora and extends onto the crystalline basement at its eastern end (Fig. 1). The other, about 12 km long and striking SW–NE, was located north of the town of Orsa and traversed only Paleozoic rocks (Fig. 1). Objectives of the profiles were to map the thickness of the sedimentary rocks and their internal structure, and if possible, identify structures in the crystal- line basement below. Results from the profiles will be used as a starting point for defining future deep drilling locations in the area.2. Geological setting The Siljan structure was formed by a bolide that impacted the terrestrial landscape of central Sweden in the late Devonian (latest Frasnian; 377 Ma, Reimold et al., 2004, 2005). The structure is con- centric with a central rise composed of Proterozoic magmatic rocks or younger Proterozoic meta-sediments and meta-volcanic rocks (western side) across a 33 km wide area in the east–west direction. The obvious circular structure outside the central rise (here called the ring) is outlined mainly by a ring of lakes and consists largely of marine Ordovician and Silurian sediments and Silurian to Devonian terrestrial sandstones, extending the east–west diameter to about 45 km (Fig. 1). It is northern Europe's largest crater, and may have played a casual role in the Fransnian/Famennian mass extinction (Reimold et al., 2004). 90 C. Juhlin et al. / Tectonophysics 580 (2012) 88–99The transient crater that formed at impact may have been anywhere between 22 and 42 km (Grieve, 1988; Holm et al., 2011; Juhlin and Pedersen, 1987, 1993; Kenkmann and von Dalwigk, 2000a). The pre- served central risemay represent the peak ring (Grieve, 1988), although Henkel and Aaro (2005) suggested that the peak ring was 20 km wide. The final diameter of the crater may have been anywhere between 52 and 90 km (Grieve, 1988; Henkel and Aaro, 2005; Holm et al., 2011; Kenkmann and von Dalwigk, 2000a,b) (Fig. 2). The ring structure itself has been denoted as an annular ring graben, a mega-block zone, or a mega-breccia (Collini, 1988; Henkel and Aaro, 2005; Juhlin and Pedersen, 1987).Within themega-blocks the Paleozoic successions are partly well preserved, albeit poorly exposed, and the stratigraphy is often continuous. It is clear that the mega-blocks now preserved in the ring could not have been within the transient crater, but rather represent radial inflow of the collapsed rim of the transient crater in concert with mass flow from the central area (Henkel and Aaro, 2005). A large transient crater would imply significant horizontal displacements of the mega-blocks (6–7 km; Kenkmann and von Dalwigk, 2000b), while a smaller crater would imply primarily vertical movements of blocks through localized sagging (listric faulting), which is what the reflection seismic data show in this paper, as well as in Juhlin and Pedersen (1987, 1993).Lake Ore Lake Orsa Lake Siljan Rättvik 0 10 km Stumsnäs 1 Lake re Lake Siljan 52 km A Stumsnäs 1 Lake Ore Lake Orsa Lake Siljan 75 km C Fig. 2. Crater models developed for the Siljan impact structure. The entire circles denote the final crater. The center of the crater is based on estimates by Juhlin and Pedersen (1987). ( Estimates by Grieve (1988) and Juhlin and Pedersen (1993) based on seismic studies. (b) A (2000a) who delimited the size from distribution of fractures in the crystalline bedrock in showing the estimated peak ring. Based on geomorphological data by Henkel and Aaro ( size of 90 km. Based on the 32–38 km estimate of the transient crater and the maximum fi features.Even though the reflection seismic data indicate a relatively simple structure, at the scale of the resolution of the data, there are geological indications of more complex structures in the area. Outcrops in the Siljan Ring are few, but the heavily tectonized nature of the sedimentary successions is often evident, generally with sharply inclined or nearly overturned packages of crystalline basement and/or sediments. One of the more important sedimentary features in the area is the large Late Ordovician Boda Limestonemudmounds, which are quarried commer- cially. Several of these are steeply inclined and all are heavily fractured (mega-brecciated; Henkel and Aaro, 2005) as a result of immediate post-impact disturbances. Recent investigations of new drill cores from the Siljan Ring have dra- matically changed the established stratigraphic view of the successions within the ring (Lehnert et al., 2012). Instead of the classical Ordovician carbonate succession separated by a stratigraphic gap from the Silurian shale dominated sequence, it is now clear that several facies belts exist. During the Middle–Late Ordovician there was a carbonate platform suc- cession towards the present day east, separated by a backbulge basin from the peripheral to the present day west in the Mora area. In Mora an erosional unconformity is observed, showing a considerable hiatus spanning the early Darriwilian to the upper Llandovery (preliminary dating; Lehnert et al., 2012). About 224 m of Silurian strata overlayLake Ore Lake Orsa Lake Siljan Rättvik Stumsnäs 1 Mora 1 Lake re Lake Siljan 65 km B Stumsnäs 1 Lake Ore Lake Orsa Lake Siljan 90 km D diameter of the transient crater, while the stippled circle denotes the diameter of the a) A final crater measuring 52 km and pre-erosion transient crater measuring 26 km. 42 km transient crater and a final size of 65 km. Based on Kenkmann and von Dalwigk terpreted as impact‐related. (c) A final crater of at least 75 km, with the inner ring he 2005). (d) A transient crater with a pre-erosion mean diameter of 35 km and a final nal diameter suggested by Holm et al. (2011), from distribution of shock metamorphic Table 1 Data acquisition. Mora Orsa Spread type Split Number of channels 320 (160–160) Near offset 0 m Geophone spacing 10 m Geophone type 28 Hz single Source spacing 20 m Source type VIBSIST 3000 Hit interval between hammer blows 100–200 ms Sweeps per source point 3–4 Nominal fold 80 Recording instrument SERCEL 428UL Sample rate 1 ms Field low cut Out Field high cut 400 Hz Record length 30 s Profile length 10 km 12 km Source points 391 542 Dates acquired 3/6–9/6, 2011 10/6–15/6, 2011 91C. Juhlin et al. / Tectonophysics 580 (2012) 88–99here a thin Lower–Middle Ordovician carbonate succession (about 21 m) in theMora 001 core (Fig. 1), drilled and owned by Igrene AB. There is no evident contact to the Orsa Sandstone or equivalent strata as is consistent with the partly coeval Nederberga Formation in the Skattungbyn area in the northern part of the ring structure (Fig. 1). The age of the uppermost preserved strata exposed at Mora is yet to be determined. Earlier acquired reflection seismic data show approximately 350 m of Paleozoic sediments preserved above the crystalline basement in the eastern part of the ring (Collini, 1988), and Thorslund (in Rondot, 1975) estimated a total of 400–500 m of these sediments to be present. Where exposed, the sedimentary successions startwith the Tremadocian, although Cambrian sediments may also be present in the area (Ebbestad and Högström, 2007). There are no continuous sections in the Siljan area, so the stratigraphy has been pieced together froma few key sections such as Kårgärde near the town of Orsa, Fjäcka on the northeast side of the ring, Amtjärn near the town of Rättvik, Styggforsen near Boda Church and the new Igrene drill cores (see Fig. 1). Exposures of the Ordovician– Silurian boundary are seen in a few quarries where the large Boda Lime- stone mud mounds are exposed. The development of strata in between the larger mounds and the Silurian succession has not been well under- stood, and the transition to the terrestrial Silurian/Devonian Orsa Sand- stone is only seen in a few drill cores (Petalas, 1985). As a consequence, the total thickness of the Paleozoic succession in the Siljan Ring is not known and, with respect to distinct facies belts, may have been variable depending on the different paleo-environments. Regardless, the Siljan area has become the standard reference area for a number of Ordovician and Silurian type formations, units and/or biozones in Sweden (see Bergström, 2007). Rough calculations of sediment thickness usingmeasurements quot- ed for each unit in Ebbestad and Högström (2007) and Lehnert et al. (in press) suggest that the classic Ordovician succession is at the minimum 115 m thick and more than 133 m thick, albeit here excluding the ~35 m thick Kullsberg mounds and the ~100 m thick Boda mounds which formed during Late Ordovician times on the outer platform. Adding the thickness of the Boda Limestone (e.g. thickness of all beds from the Obolus beds to the Fjäcka Shale plus Boda Limestone), gives an estimate of between 185 and 215 m for the Ordovician sediments. In the Mora 001 core, the Silurian succession is at least 224 m thick (Lehnert et al., 2012), although it does not reflect the depositional thick- ness over the entire Siljan area. The thickest succession of Orsa Sand- stone known is at least 58.9 m (Petalas, 1985). Adding up these highly uncertain estimates gives a minimum thickness of the Paleozoic sedi- ments in the order of 460 to 500 m. 3. Data acquisition Seismic data (Table 1) were acquired along two profiles in the west- ern part of the Siljan Ring area in the summer of 2011 (Fig. 1). Since ob- jectives were to both image the Paleozoic sedimentary rocks and the deeper crystalline basement (down to 3 km) along longer profiles a compromise between station spacing and profile length was necessary. In addition, it was expected that signal penetration could be an issue at some locations due to large amounts of sand in the near-surface. There- fore, a relatively powerful sourcewith a source spacing of 20 m and a re- ceiver spacing of 10 m was employed. If the Paleozoic rocks extend down to 500 m then this source and receiver spacing corresponds to an effective fold of about 25 at this depth with a normal moveout (NMO) stretch of about 50%. At 250 m depth the effective fold is half of this. This lower fold at shallower depth implies that degraded images in those parts of the final stacked sectionwhere the source couplingwas poor may be expected. Datawere first acquired along theMora profile during a 6 day period in the beginning of June 2011 using a tractor mountedmechanical ham- mer (see Juhlin et al. 2010 for details on the source). Acquisition started on the eastern end of the profile on granitic bedrock and proceeded westwards. It was not possible to lay cables across the outlet to LakeOrsa (stations 1016–1035 in Fig. 3) so the acquisition geometry had to be adjusted in this area. A cabled fixed spreadwas used east of the outlet with 36 wireless units west of the outlet while the source was active on the eastern part.When all possible source points had been activated east of the outlet the cabled spread was moved to the west of the outlet and the 36 wireless units to the east of it. This strategy allowed us to acquire data to fill the gap below the outlet without having to repeat source points and avoided us having to transport the source back and forth through the town of Mora. In addition to the logistics with the outlet, there were many locations along the Mora profile where the source could not be activated due to the near-surface being too soft or that the profile ran too close to residential areas. These locationsweremostly on the eastern half of theprofile (Fig. 3). Surface deposits east of the Lake Orsa outlet consist mainly of moraine while west of the outlet the sur- face deposits are mainly glacial and fluvial sands. After completing the Mora profile the spreadwas moved to the Orsa profile (Fig. 4) and acquisition continued using the same mechanical source. Logistics were less problematic along the Orsa profile, but there were some locations in the southern part of the profile where it was not possible to activate the source (Fig. 4). South of station 350 sur- face deposits are mainly fluvial sands while to the north of it glacial deposits dominate. In addition to the cabled spread that was rolled along during the acquisition, twelve 3-component wireless units were placed out along the profile at a spacing of 1 km. These units recorded nearly all of the source points during the acquisition, but presentation of data from these units is not the focus of this paper. However, these data may be useful in the future for building velocity models based on refracted waves. 4. Data processing Datawere decoded using a shift and stackmethod based on Park et al. (1996). Data quality varied significantly along the profiles depending upon the near surface conditions. The thickness of the uppermost loose deposits varies considerably in the area, as does the groundwater level. In- terestingly, the depth to the groundwater level appeared to have less influence on the data quality than may have been expected. First break quality is nearly the same regardless of how much dry sand is present (Fig. 5). The thickness of the groundwater bearing loose sediments appears to be a more important factor influencing data quality. Where these loose sediments are thick a very ‘ringy’ signal after the first arrival is present. This signal masks weaker reflections and multiple reflections are generated that arrive within a cone that is limited by a velocity of about 1700 m/s. Within this cone it is difficult, if not impossible, to trace reflections with the receiver spacing used. Any reflections arriving at late times will be masked by this cone. When this groundwater-bearing 470 480 6760 6770 W-E (km) N -S (k m) Silurian clastics Ordovician carbonates Neoproterozoic dolerites Precambrian felsic intrusives Precambrian felsic extrusives Precambrian sedimentary rocks Fig. 3. Mora profile location. Blue line indicates where the geophones were placed. When filled with red then the source was activated at those locations along the profile. Yellow line shows the CDP processing line that the data were stacked along. Geological background map provided by the Geological Survey of Sweden (www.sgu.se). 92 C. Juhlin et al. / Tectonophysics 580 (2012) 88–99layer of loose sediments is thinner or absent this cone is less prevalent (Fig. 5a, left side of spread) and reflections are more easily observed in raw source gathers. The thickness of this problematic layer varies consid- erably along theMora profile, while along theOrsa profile this layer is less of a problem and appears to be presentmainly in the southern part of the profilewherefluvial sediments are present at the surface (Fig. 6a). Further north along the Orsa profile the data quality is generally good with clear first arrivals across the entire spread (Fig. 6b). Note how the apparent ve- locity increases at the further offsets in Figs. 5a and 6b, indicating higher velocity rock at depth. The variable data quality along theprofiles, especially theMoraprofile, alongwith variable velocities, both vertically and horizontally, was a pro- cessing challenge. It is difficult to estimate the velocity in the dry loose de- posits since the direct wave through them is only recorded on the closest geophone or two, but based on these arrivals a velocity of 600 m/s is rea- sonable for these deposits. When water saturated, these loose deposits have a velocity on the order of 1700 m/s. The first main refractor that is observed has a velocity of about 3000–3200 m/s, corresponding to the uppermost Paleozoic rocks. A deeper refractor is sometimes observed with a velocity on the order of 4000–4200 m/s that probably represents a deeper succession of Paleozoic rocks. East of the Lake Orsa outlet on the Mora profile, refractor velocities of about 5000 m/s are observed. Bothprofiles followed roads and forest tracks, resulting in crooked line data being acquired. One option for handling such data is to define a sla- lom stacking line that follows the general trend of the mid-points. How- ever, this strategy can result in false structures being introduced into the final stacks where the profile changes direction rapidly (e.g. Wu etal., 1995). Instead, straight stacking lines were used with Common Data Point (CDP) bins defined perpendicular to the stacking line with the true acquisition geometry still being honored. That is, it was not assumed that the data had been acquired along straight lines. Since the stacking (or CDP) lines are straight they can be shifted at will in the direction normal to them and the same result will be obtained in the processing. Therefore, the location of the stacking lines in Figs. 3 and 4 is only schematic, and they may be shifted as explained above without affecting the processing results. Mostly standard processing steps were applied to the data after CDP binning. Due to that most of the reflections are observed best at wide angles a large normal NMO stretch was allowed. This large stretch re- duces the apparent frequency content of the data in the stacked sec- tions. However, if the maximum stretch was decreased to a more normal value of 40% then some reflections were completely muted in the source gathers and no additional reflections appeared in the stacks. Increasing the allowable stretch to above 70% did not only increase am- plitudes in the stacked sections, but also decreased the frequency con- tent even more. A stretch value of 70% in the NMO processing step therefore seemed to be a reasonable compromise. Due to the often ‘ringy’ nature of the data in the Mora profile and the more prevalent noise cone, both a top mute and a bottom mute were applied before stacking. Top muting was to ensure that first arrivals were not stacked in as reflected energy. Bottom muting was applied to reduce the influ- ence of the noise cone in the uppermost 300 ms of the stacked section. A complete list of processing steps for the two profiles is provided in Table 2. Silurian clastics Ordovician carbonates Neoproterozoic dolerites Precambrian felsic intrusives Precambrian felsic extrusives Precambrian sedimentary rocks W-E (km) N -S (k m) 480 6780 Fig. 4. Orsa profile location. Blue line indicates where the geophones were placed. When filled with red then the source was activated at those locations along the profile. Yellow line shows the CDP processing line that the data were stacked along. Geological background map provided by the Geological Survey of Sweden (www.sgu.se). 93C. Juhlin et al. / Tectonophysics 580 (2012) 88–995. Results Stacked and migrated sections are shown for the Mora and Orsa profiles in Figs. 7 and 8, respectively. The Mora profile is characterized by reflective packages of variable dip with the only longer clear sub-horizontal reflection being present in the west between CDPs 350 and 700 at an approximate time of 200 ms (c. 250 m). The rest of the profile either shows a transparent seismic response, reflections with either a predominantly easterly or westerly dip, or shorter packages of sub-horizontal reflections (Fig. 7). There are no clear reflections from below 500 ms. Note the generally higher frequency content in the data east of CDP 1500 where the bed- rock consists of Ordovician carbonate rocks and Precambrian crystalline rocks. Dipping events in the upper 400 ms, with opposite dips, between CDPs 500 and 700 may be from real steeply dipping structures since they have apparent velocities of about 3500 m/s. However, they may also result from direct or surfacewaves that have not been removed en- tirely during the processing. The more moderately, mainly east dipping events, between CDPs 800 and 1000 probably represent a sub-surface structure that has been significantly disturbed. The Orsa profile shows a more consistent image with mainly sub- horizontal reflections present in the uppermost 400 ms throughout the profile (Fig. 8). Only in the south are transparent zones present, be- tween CDPs 100 to 300 and 550 to 650. Dipping reflections are most ap- parent between CDPs 300 and 500 in the south and in the northernmostpart of the profile around CDP 2200. A deeper sub-horizontal reflective zone is present in the central parts of the profile between 800 and 900 ms. There are also indications of dipping reflections above this sub-horizontal reflectivity. Evidence for faulting is present throughout most of the section. Faults with minor throws are obvious between CDPs 1250 and 1700 in the upper 200 ms. Further north, between CDPs 1750 and 2000, a graben-like structure is present in the upper 400 ms containing faults with throws on the order of 100 ms (200 m). 6. Seismic modeling Based on the refractions observed in the source gathers and the pres- ence of wide-angle reflections there are clear velocity contrasts in the sub-surface along the profiles that can produce reflections. However, the stacked and migrated images show a complex and sometimes spo- radic reflectivity pattern that is difficult to interpret based on geometry alone. It is not even clear if the crystalline Precambrian basement has been imaged along the profiles when only inspecting the stacked sec- tions. To aid in the interpretation, CDP super gathers of the raw data were generated to allow velocity as a function of depth (time) to be estimated along the profiles (Fig. 9). Data from five adjacent CDPs were binned at an offset interval of 10 m and then stacked within com- mon offset bins so that only one trace is presented for each offset in the super gathers. Onmost super gathers there are clear breaks in the slope of the first arrivals with the slope decreasing with offset, indicating the 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Ti m e (s) 700 750 800 850 900 950 1000 1050 Station SP842 300 350 400 450 500 550 600 Station SP460 Fig. 5. (left panel) Source gather from SP 842 on the Mora profile (Fig. 3) showing a relatively small delay in the zero crossing of the main first break energy which indicates that the slow dry near-surface loose deposits are relatively thin at this source location. (right panel) Source gather from SP 460 on the Mora profile (Fig. 3) showing a longer delay in the zero crossing, a ringy signal after the first arrival and a cone multiple noise. The latter two characteristics of the data are probably due to a relatively thick groundwater layer at this location which energy gets trapped in. 94 C. Juhlin et al. / Tectonophysics 580 (2012) 88–99arrival of refractions from deeper layers. The arrival times were then modeled for every 50th CDP assuming a horizontally layered sub- surface and increasing velocity with depth to generate velocity-depth models along the profiles. The assumption of a horizontally layered sub-surface is true along much of the profiles, but is obviously violated on some parts. Dipping layers and rapid lateral variations in velocity can significantly perturb the arrival times compared to the 1D assumption.0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Ti m e (s) 50 100 150 200 Station SP30 400 Fig. 6. (left panel) Source gather from SP 30 on the Orsa profile (Fig. 4) showing a moderat arrivals and some signs of a noise cone due to trapped energy in the groundwater layer. (ri sponse from the more northern part of the profile. Here, the ringyness is less prevalent andHowever, the velocity-depth models do show general consistency along the profiles and aid in the interpretation of the reflection seismic images. As a further check on the modeling, synthetic seismograms were generated based on the obtained velocity-depth models using an elastic modeling program (Juhlin, 1995). Poisson's ratio was assumed to be on the order of 0.33 and the density increasing with increasing450 500 550 600 650 700 Station SP548 e delay in the zero crossing of the main first break energy, a ‘ringy’ signal after the first ght panel) Source gather from SP 460 on the Orsa profile (Fig. 4) showing a typical re- the multiple noise cone is generally absent. Table 2 Data processing. Step Parameters Mora Orsa 1 Read decoded VIBSIST data 2 Bulk static shift to zero time 3 Apply geometry 4 Pick first breaks 5 Spherical divergence correction 6 Trace editing 7 Trace balance: 0–3000 ms 8 Spectral equalization: 0–600 ms: 50–80–200–240 Hz 700–1500 ms: 40–70–180–240 Hz 9 Time variant bandpass filter: 0–400 ms: 50–80–240–360 Hz 450–600 ms: 45–70–210–300 Hz 700–1000 ms: 40–60–180–270 Hz 1100–3000 ms: 35–50–150–225 Hz 10 Refraction statics: datum 160 m, replacement velocity 3000 m/s Refraction statics: datum 180 m, replacement velocity 3000 m/s 11 Residual statics 12 Median filter: 11 traces, 3 samples, 5300 m/s, subtract Median filter: 11 traces, 3 samples, 1600, 2000, 2400 m/s, subtract 13 AGC: 50 ms 14 Mute: top and bottom No mute 15 Residual statics 16 Velocity analysis 17 NMO correction: 70% stretch mute 18 Trace balance 19 FX Decon: 19 trace window 20 No dip filter Dip filter 1.5 ms/trace cutoff 21 Trace balance 22 Stolt migration: 200–3000, 1000–4000 ms/s 95C. Juhlin et al. / Tectonophysics 580 (2012) 88–99P-wave velocity for the modeling. Fig. 9a shows a CDP super gather at CDP 600 on the Mora profile (Fig. 7), a location where the structure appears to be relatively sub-horizontal. The first arrivals calculated0 2 4 Distanc 0 1 Ti m e (s) Ti m e (s) 200 400 600 800 100 CD W 0 2 4 Distan 0 1 200 400 600 800 10 CD W a b Fig. 7. Mora profile showing the (a) stfrom the velocity-depth model (Table 3) match well with the observed first arrivals (Fig. 9a). Note that the low velocity unsaturated loose de- posits (layer 1 in Table 3) appear to be relatively thick here (17 m), but that signals are still clear out to 1500 m offsets in the real data. Most of the general characteristics of the modeled data agree well with observations in the real data. In particular, the strong multiples, both reflected and refracted, and the noise cone due to the water satu- rated loose sediments (layer 2 in Table 3) are similar in the two sections (Fig. 9a). One discrepancy is the reflected wave from the base of layer 3 that arrives earlier in themodeled data than in the real data. The reason for this is not clear, but could be related to anisotropy, with propagation velocities faster in the horizontal direction than in the vertical direction. Further west, at CDP 300 (Fig. 9b), the first arrivals are less clear at far offsets in spite of low velocity layer 1 being thinner (Table 3), implying that other factors are affecting the data quality than the uppermost dry loose sediments. First arrivals are modeled well by a 5 layer model, but there is no clear reflection from the base of layer 4 (the equivalent to layer 3 at CDP 600). This may be due to the interface being disturbed at this location or generally poorer data quality in this area. Reflections at this location are also absent on the stacked section (Fig. 7). Note that the noise cone is not modeled as well here as at CDP 600. Further east, at CDP 1300 (Fig. 9c), first arrivals from the equivalent of layer 3 at CDP 600 and layer 4 at CDP 300 are absent. Instead first arrivals from a layer with a slope corresponding to 1550 m/s are present to offsets of al- most 500 m. A refracted wave with an apparent velocity of 5100 m/s then arrives first out to about 800 m. At greater offsets it is more difficult to identify the first arrival, but the apparent velocity increases up to about 7500 m/s, a very high value. Significant multiple energy is also present in the synthetic data as is a noise cone that corresponds well in time with the real data. The reflection from the base of layer 2 arrives within this noise. Therefore, it is unlikely that this reflection would be imaged on the real data, although signs of it are present. The velocity-depth models that were determined along both pro- files have been plotted onto the stacked migrated sections to aid in6 8 e (km) 0 1200 1400 1600 1800 P E 6 8 ce (km) 00 1200 1400 1600 1800 P E acked and (b) migrated sections. 0 2 4 6 8 10 Distance (km) 0 a b 1 Ti m e (s) Ti m e (s) 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 CDP SW NE 0 2 4 6 8 10 Distance (km) 0 1 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 CDP SW NE Fig. 8. Orsa profile showing the (a) stacked and (b) migrated sections. 96 C. Juhlin et al. / Tectonophysics 580 (2012) 88–99the interpretation (Figs. 10 and 11). Prior to plotting, the depths were converted to time and a static shift corresponding to the refraction static correction at each CDP location was applied. This allows a rough com- parison to be made between the refraction models and the reflection images. The refraction models and the interpretation of the reflections images are discussed in the next section. 7. Discussion The reflection seismic images do not show a simple layered structure along the profiles, indicating that complex geological conditions are present in some parts. Interpretation of the profiles based solely on the reflection images is difficult. However, themodeled refractors aid signif- icantly in the interpretation, especially if lithologies can be assigned to the velocities. Unfortunately, no velocity logs are available through the sedimentary successions from any of the boreholes. Therefore, we have assumed the following in our interpretation: velocities less than 2100 m/s correspond to glacial sediments, velocities in the range of 2500–3200 m/s correspond to Silurian clastic rocks, velocities in the range of 3800–4200 m/s correspond to Ordovician limestones, veloci- ties in the range of 4500–5500 m/s correspond to highly fractured crys- talline rock or Precambrian meta-sedimentary/volcanic rocks and velocities greater than 5500 m/s correspond to mainly intact crystalline bedrock. In the Mora section (Fig. 10), the refractor velocity corresponding to the interpreted top of the Silurian is relatively consistent along the pro- file at about 3000–3200 m/s. This refractor is only present in the south- ern and northern parts of the Orsa profile (Fig. 11), indicating the Silurian to be thinner here, or perhaps absent in the central part. The interpreted refraction velocity at the top of the Ordovician is generallyFig. 9. CDP super gathers with corresponding elastic modeling at CDP locations (a) 600, (b) calculated from the velocity-depth models in Table 3 while the green line represents the refl 300 and the base of layer 2 for CDP 1300. Data area plotted with at reduced times with thehigher and more variable in the Orsa profile than in the Mora profile. In addition, the velocities below the base of the interpreted Ordovician on the Orsa profile are high at about 6000 m/s, indicating intact crystal- line basement is present. On the Mora profile, first arrivals with such a high apparent velocity are not observed, except at CDP 1300, suggesting that the crystalline basement is deeper and has not been mapped. At around CDP 1300 (Fig. 9c) higher apparent velocities are observed, but this is very local. The 5100 m/s velocities at CDP 1300 are probably too low to represent intact crystalline basement and the 7500 m/s velocity may be due to geometrical effects or to a high velocity mafic intrusion at depth. The deeper reflections corresponding to this high velocity boundary are also a sign of a strong impedance contrast at depth (Fig. 10). The only clear reflection from within the basement is observed on the Orsa profile, a gently dipping event at about 900 ms (Fig. 11). Given that sub-horizontal reflections further east at similar depths are generated by thin dolerite sills (e.g. Juhlin, 1990), it is likely that this reflection also is generated from a dolerite sill at about 2 km depth. However, it is also possible that it represents a detachment surface on which mega-blocks have slid. Therefore, it is of interest to drill through the interface generating this reflection. Recent studies of new drill cores in the Siljan Ring (Lehnert et al., 2012) have shown that the classical distribution in the Siljan area of an Ordovician carbonate platform followed by Silurian shales and mud- stone and the Silurian/Devonian terrestrial Orsa Sandstone has to be re-evaluated and ancient facies belts have to be reconstructed in detail for different time slices. The development of the Lower–Middle Ordovi- cian carbonate succession, from the contact to the Precambrian base- ment to the upper part of an overlying limestone (Holen Formation) is approximately uniform across the Siljan district at about 20 m. In the Mora area this limestone exhibits exposure plus erosion, interpreted as300 and (c) CDP 1300 along the Mora profile. Blue lines represent the first break times ection time corresponding to the base of layer 3 for CDP 600, the base of layer 4 for CDP reduction velocity given by RVEL. 0.0 a b c 0.1 0.2 0.3 Ti m e (s) 0.0 0.1 0.2 0.3 Ti m e (s) 0.0 0.1 0.2 0.3 Ti m e (s) 0 200 400 600 800 1000 1200 1400 Offset (m) 0 200 400 600 800 1000 1200 1400 Offset (m) 0 200 400 600 800 1000 1200 1400 Offset (m) 0 200 400 600 800 1000 1200 1400 Offset (m) 0 200 400 600 800 1000 1200 1400 Offset (m) 0 200 400 600 800 1000 1200 1400 Offset (m) CDP=300RVEL=3000.0 CDP=600RVEL=3000.0 CDP=1300RVEL=3000.0 97C. Juhlin et al. / Tectonophysics 580 (2012) 88–99 Table 3 Modeling parameters used for Fig. 9. HS refers to a half-space. Layer CDP 300 CDP 600 CDP 1300 Base depth (m) Vp (m/s) Base depth (m) Vp (m/s) Base depth (m) Vp (m/s) 1 8 600 17 600 5 600 2 27 1600 50 1700 179 1550 3 120 2520 254 3200 402 5100 4 217 3200 HS 4050 HS 7500 5 HS 4100 Fig. 10. Interpreted Mora profile. Times to refractors are plotted as short red lines at those CDP locations where a reasonable time-offset curve for the first arrivals could be picked. The velocity of the refracting interface is shown below each red line. Shown also are interpretations of the boundaries of the different geological units based on the reflection seismic images and the refraction interpretation. Yellow—base of the water saturated loose sediments, green—base of the Silurian, blue—base of the Ordovician, black—dolerite intrusion, gray—fault. 98 C. Juhlin et al. / Tectonophysics 580 (2012) 88–99due to the formation of a peripheral bulge in the Caledonian foreland basin. Later, the limestonewas covered by amore than 200 m thick Silu- rian clastic succession. This Silurian shale basin in the Mora area (prob- ably thinning or partially eroded towards the east) may partly explain why the basement is apparently deeper along the Mora seismic profile than along the Orsa profile. The only location along the Mora profile that the reflection seismic data indicate a relatively undisturbed layering of the sedimentary successions is between CDPs 350 and 700. Here the interpreted top of the Ordovician corresponds to a clear seismic reflection, indicating little internal deformation, and the refraction modeling indicates the Silurian to be about 200 m thick below CDP 600 (Table 3). If the deeper reflection at 0.4 s (Fig. 10) corresponds to the top of the base- ment then the unit below is about 400 m thick, giving a total thickness of the Paleozoic successions of about 600 m here. Whether the lower- most unit consists entirely of Ordovician limestones or whether there may be Cambrian units present at depth cannot be determined from the seismic data. On the rest of theMora profile there are only short seg- ments where reflections directly correlate with the results from the refraction interpretation, indicating that the sedimentary successions are heavily disturbed. The depth to the crystalline basement is relatively well determined along the Orsa profile, generally being on the order of 200–300 m, except south of CDP 500 and in the heavily faulted northern part of the profile (CDPs 1750 to 2000) that was mentioned earlier. This is significantly less than along theMora profile, suggesting that the facies may be different north of Lake Orsa compared to that found in the Mora area. Intense thrusting during the emplacement of megablocks is observed in the yet to be described Stumsnäs 1 (Fig. 1) core of Igrene AB (repeated thrusting of granitic basement over fault-bounded Ordovician limestone units) and the Nittsjö trench (Fig. 1) (thrust slices of Lower/Middle Ordo- vicianmarls in structural contactwith Proterozoicmeta-volcanoclastics). In addition, the occurrence of a Silurian clastic succession (NederbergaFormation), a facies typical of the Silurian shale basin in the western part of the impact structure, in the north-central part of the ring structure at Skattungbyn reflects also some large-scale overthrusting and not sim- ple sliding of blocks from the rim of the crater towards the central uplift. Similar thrusting may be expected to be present in the Mora area and may explain the lack of coherency of reflections along most of the Mora profile. 8. Conclusions The reflection seismic data presented here shed new light on the Paleozoic successions in the western Siljan Ring. They show a complex and disturbed reflection pattern, especially along the Mora profile. Re- flections are more coherent on the northern Orsa profile. Interpretation of the reflection images was greatly aided by combining information from refraction modeling, allowing depths to the top of Ordovician limestone to be estimated along both profiles. New core data from a borehole drilled in the town of Mora shows a thick Silurian shale sequence (more than 200 m) above the Ordovician limestone in this part of the Siljan Ring. Where reflections are clear from the interpreted base of the Silurian on the Mora seismic profile (CDPs 350 to 700) we estimate the Silurian to be about 200 m thick, in- dicating that this shale sequence extends at least this far to the north. Depth to the crystalline bedrock is generally greater along theMora pro- file than theOrsa profile, with the Paleozoic successions interpreted to be as thick as 600 m in thewestern part of theMora profile. Both the reflec- tion images and the refraction interpretation indicate the bedrock to be highly faulted. In the northern part of the Orsa profile the data indicate normal faultswith throws of up to 200 m. Significant faulting of the bed- rock is also interpreted in the eastern part of the sedimentary successions on the Mora profile. In order to study the distribution of significant sedimentary changes (limestones vs. shales vs. sandstone units) and to reconstruct the Fig. 11. Interpreted Orsa profile. Times to refractors are plotted as short red lines at those CDP locations where a reasonable time-offset curve for the first arrivals could be picked. The velocity of the refracting interface is shown below each red line. Shown also are interpretations of the boundaries of the different geological units based on the reflection seismic images and the refraction interpretation. Yellow—base of the water saturated loose sediments, green—base of the Silurian, blue—base of the Ordovician, black—dolerite intrusion, gray—fault, pink area—possible reflective zone in the Precambrian basement. 99C. Juhlin et al. / Tectonophysics 580 (2012) 88–99extension of coeval facies belts, the position of megablocks and thrust slices, higher resolution shallow seismic studies together with new drill cores are required. One new core must include the Silurian siliciclastic succession in the Orsa area, where lithostratigraphic control is needed for seismic interpretation of the data reported here. Future higher resolution shallow seismic studies will help to document the contacts between various Silurian sandstone and shale units as well as between Silurian siliciclastics and underlying Ordovician shelf carbon- ates resting on Precambrian basement rocks. Both additional seismic data and shallow drilling are required before the location of a deeper borehole can be finalized. Acknowledgments The Swedish Research Council (VR) funded the reflection seismic data acquisition (project number 2009‐4492) and is gratefully acknowledged. GLOBE Claritas™ under license from the Institute of Geological and Nuclear Sciences Limited, Lower Hutt, New Zealand was used to process the seismic data. References Bergström, S.M., 2007. The Ordovician conodont biostratigraphy in the Siljan region, south- central Sweden: a brief review of an international reference standard. In: Ebbestad, J.O.R., Wickström, L.M., Högström, A.E.S. (Eds.), WOGOGOB 2007. 9th Meeting of the Working Group on Ordovician Geology of Baltoscandia. Field Guide and Abstracts: Sveriges Geologiska Undersökning Rapporter och Meddelanden, 128, pp. 26–41. 63–78. Bodén, A., Erikson, K.G., 1988. Deep Drilling in Crystalline Bedrock; Vol. 1: The Deep Gas Drilling in the Siljan Impact Structure, Sweden and Astroblemes. Springer Verlag, Berlin . 364 pp. Collini, B., 1988. Geological Setting of the Siljan Ring Structure Deep Drilling in Crystal- line Bedrock, Vol. 1: The Deep Gas Drilling in the Siljan Impact Structure, Sweden and Astroblemes. Springer Verlag, Berlin, pp. 349–354. Ebbestad, J.O.R., Högström, A.E.S., 2007. Ordovician of the Siljan district, Sweden. In: Ebbestad, J.O.R., Wickström, L.M., Högström, A.E.S. (Eds.), WOGOGOB 2007. 9th meeting of the Working Group on Ordovician Geology of Baltoscandia. Field guide and Abstracts: SGU Rapporter och meddelanden, 128, pp. 7–26. Grieve, R.A.F., 1988. The formation of large impact structures and constraints on the nature of Siljan. In: Boden, A., Eriksson, K.G. (Eds.), DeepDrilling in Crystalline Bedrock. : TheDeep Gas Drilling in the Siljan Impact Structure, Sweden and Astroblemes, Proceedings of the International Symposium, vol. 1. Springer Verlag, Berlin, pp. 328–348. Henkel, H., Aaro, S., 2005. Geophysical investigations of the Siljan impact structure—a short review. In: Koeberl, Henkel, H. (Eds.), Impact Tectonics. Springer Verlag, Berlin, pp. 247–283. Högström, A.E.S., Sturkell, E., Ebbestad, J.O.R., Lindström, M., Ormö, J., 2010. Concentric impact structures in the Palaeozoic of Sweden—the Lockne and Siljan craters. GFF 132, 65–70.Holm, S., Alwmark, C., Alvarez,W., Schmitz, B., 2011. Shock barometry of the Siljan impact structure, Sweden. Meteoritics and Planetary Science 46, 1888–1909. Juhlin, C., 1990. Interpretation of the reflections in the Siljan Ring area based on results from the Gravberg-I borehole. Tectonophysics 173, 345–360. Juhlin, C., 1995. Finite difference elastic wave propagation in 2-D heterogeneous trans- versely isotropic media. Geophysical Prospecting 43, 843–858. Juhlin, C., Pedersen, L.B., 1987. Reflection seismic investigations of the Siljan impact structure, Sweden. Journal of Geophysical Research 92, 14113–14122. Juhlin, C., Pedersen, L.B., 1993. Further constraints on the formation of the Siljan impact crater from seismic reflection studies. Geologiska Föreningens i Stockholm Förhandlingar 115, 151–158. Juhlin, C., Dehghannejad, M., Lund, B., Malehmir, A., Pratt, G., 2010. Reflection seismic imaging of the end-glacial Pärvie Fault system, northern Sweden. Journal of Ap- plied Geophysics 70, 307–316. Juhojuntti, N., Juhlin, C., 1998. Seismic lower crustal reflectivity and signal penetration in the Siljan Ring area, Central Sweden. Tectonophysics 288, 17–30. Kenkmann, T., von Dalwigk, I., 2000a. Centro‐symmetrical material flow during crater modification: structural implications. Lunar and Planetary Science Conference XXXI, Abstract 1041. Kenkmann, T., vonDalwigk, I., 2000b. Radial transpression ridges: a new structural feature of complex impact craters. Meteoritics and Planetary Science 35, 1189–1202. Lehnert, O., Meinhold, G., Bergström, S.M., Calner, M., Ebbestad, J.O.R., Egenhoff, S., 597 Frisk, Å.M., Hannah, J.L., Högström, A.E.S., Huff, W.D., Juhlin, C., Maletz, J., Stein, H.J., Sturkell, E., Vandenbroucke, T.R.A., 2012. New Ordovician-Silurian drill cores from the Siljan impact structure in central Sweden - an integral part of the Swedish Deep Drilling Program. GFF 134, 87–98. http://dx.doi.org/10.1080/11035897. 2012.692707. Lorenz, H., 2010. The Swedish Deep Drilling Program: for science and society. GFF 132, 25–27. Papasikas, N., Juhlin, C., 1997. Interpretation of reflections from the central part of the Siljan Ring impact structure based on results from the Stenberg-1 borehole. Tectonophysics 269, 237–245. Park, C.B., Miller, R.D., Steeples, D.W., Black, R.A., 1996. Swept impact seismic technique (SIST). Geophysics 61, 1789–1803. Petalas, C., 1985. Sedimentary petrology of the Orsa sandstone, central Sweden. UUDMP research report, 40, pp. 1–138. Reimold,W.U., Kelley, S.P., Sherlock, S., Henkel, H., Koeberl, C., 2004. Laser argon dating of melt breccias from the Siljan impact structure, Sweden—implications for possible relations to late Devonian Extinction events. Lunar and Planetary Science Conference, XXXV, Abstract 1480. Reimold, U., Kelley, S.P., Sherlock, S., Henkel, H., Koeberl, C., 2005. Laser argon dating of melt breccias from the Siljan impact structure, Sweden: implications for a possible relationship to Late Devonian extinction events. Meteoritics and Planetary Science 40, 591–607. Rondot, J., 1975. Comparison entre les astroblemes de Siljan, Suede et de Charlevoix, Quebec, vol. 6. Bulletin of the Geological Institute of the University of Uppsala, N.S., pp. 85–92. Wu, J., Milkereit, B., Boerner, D.E., 1995. Seismic imaging of the enigmatic Sudbury Structure. Journal of Geophysical Research 100 (B3), 4117–4130.