<<

Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology

journal homepage: www.elsevier.com/locate/palaeo

Did mangrove communities exist in the Late of the T Basin, ? ⁎ Stephen McLoughlina, , David W. Haigb, Mikael Siverssonc,e, Elisabeth Einarssond a Department of Palaeobiology, Swedish Museum of Natural History, Box 50007, S-104 05 Stockholm, Sweden b Centre for Energy Geoscience, School of Earth Sciences, The University of Western , Perth 6009, Australia c Department of Earth and Planetary Sciences, Western Australian Museum, Welshpool, WA 6106, Australia d Department of Geology, Lund University, S-223 62 Lund, Sweden e School of Molecular and Life Sciencesc, Curtin University, Kent Street, Bentley, WA 6102, Australia

ARTICLE INFO ABSTRACT

Keywords: Previous inferences of -dominated communities occupying mangrove-like depositional settings in the , Sweden, during the late early Campanian are reassessed. A significant percentage of Oysters (Acutostrea incurva) from the mammillatus zone in Bed 3 at Åsen bear indentations on their left valves indicating attachment to plant axes. Many of these axes bear morphological features characteristic of the distal subaerial portions of woody plant branches and appear to have been rafted into the marine environment rather than representing in situ mangrove stems and roots. Foraminiferal assemblages recovered from sediment Mangroves ff Sweden within the oyster body cavities di er from modern mangrove-community associations by the absence of siliceous agglutinated Foraminifera, the presence of diverse and relatively abundant , relatively common triserial Buliminida, and a notable percentage of planktonic taxa. Chondrichthyan teeth assemblages from the same beds are similarly incompatible with the interpretation of a mangrove depositional environment based on compar- isons with the distribution of related extant taxa. Apart from oyster shells and belemnite rostra, these beds are notably depauperate in diversity and abundance of macroinvertebrate remains compared with coeval carbonate shoal and rocky shoreline assemblages from the same basin. The collective palaeontological and sedimentolo- gical evidence favours an inner neritic sandy-substrate setting, but not nearshore or mangrove-like depositional environment for the oyster-rich Bed 3 at Åsen. The absence of mangrove-like assemblages at Åsen is consistent with the development of modern mangrove ecosystems much later (during the and ) based on the global palynological record.

1. Introduction environment in the Kristianstad Basin by Christensen (1975) and later authors all refer to the Åsen locality (Fig. 1). This area was situated The palaeoecology of some local marine invertebrate assemblages in within a drowned, wedge-shaped river valley, open to the sea in the lower Campanian strata of the Kristianstad Basin, southern Sweden, has south (Siversson et al., 2016). The presumed mangrove-like setting at long been interpreted in terms of mangrove or mangrove-bordering Åsen has in turn been used to infer that the environment was suitable estuarine communities fringing an archipelago (Spjeldnæs, 1975; only for opportunistic marine organisms adapted to strong fluctuations Sørensen and Surlyk, 2008). The inference apparently extends back to in salinity and water turbidity (Sørensen and Surlyk, 2008). Siversson the work of Nilsson (1827) and Christensen (1975) who noted examples et al. (2016) questioned the palaeoenvironmental interpretation of of the oyster Acutostrea incurva Nilsson, 1827 with a cylindrical groove Christensen (1975) and Sørensen and Surlyk (2008), citing the abun- on the outside of the left valve interpreted to represent attachment to dance of teeth of angel sharks (Squatina spp.). Although some of roots or detached branches (Skarby, 1986). More recently, Sørensen extant angel sharks occur in and may be relatively common in and Surlyk (2008, fig. 2) illustrated an example of an A. incurva left the vicinity of estuarine fronts (Vögler et al., 2008), there are no records valve from the Kristianstad Basin with what they interpreted to be an of them being present in abundance in close proximity to mangroves ‘attachment imprint representing growth on branches or rhizomes of (see Compagno, 1984). mangrove-like trees’. The discussions of a Campanian mangrove Since oyster fixation to woody plants is key to the interpretation of

⁎ Corresponding author. E-mail address: [email protected] (S. McLoughlin). https://doi.org/10.1016/j.palaeo.2018.03.007 Received 30 December 2017; Received in revised form 5 March 2018; Accepted 5 March 2018 Available online 14 March 2018 0031-0182/ © 2018 Elsevier B.V. All rights reserved. S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

SWEDEN 56°15'N Kristianstad Basin Skåne Hanö IVÖ KLACK ÅSEN AXELTORP Bay Hässleholm Basin Vinslöv Block RYEDAL

IGNABERGA 50 km ULLSTORP Nävlingeåsen Kristianstad N 56°00'

KRISTIANSTAD BASIN

Linderödsåsen Town

10 km HANÖ BAY Locality

Upper Cretaceous

Other bedrock 13°30'E 14°00' 14°30' 55°45'

Fig. 1. Schematic geological map of the Kristianstad Basin in Skåne, southern Sweden (after Siverson, 1992; Einarsson et al., 2016).

Campanian ecosystems at Åsen outlined by Sørensen and Surlyk (2008), 2016; Halamski et al., 2016; Siversson et al., 2016). The Campanian it is pertinent to consider the history and ecology of ostreoid bivalves floras are dominated by broad-leafed dicotyledonous angiosperms with and mangroves. Oysters originated at least by the Middle and subsidiary conifers and but taxa with mangrove-like habits have perhaps as early as the (Márquez-Aliaga et al., 2005). Early not been identified (Halamski et al., 2016). This in- forms attached to rocky or other skeletal substrates but modern oysters terval corresponds to a time of rapid diversification among angiosperms may attach to a broad range of natural and artificial materials including and their rise to dominance in the global flora (Friis et al., 2011). shell, rock, glass, wood, ropes and even plastics (Nelson, 1924; Barry, Within the marine invertebrate assemblage, numerous fossils of Acu- 1981; Wallace, 2001; Gardunho et al., 2012), on which they commonly tostrea incurva bear clear indications of attachment to elongate, more- form gregarious communities. Hard and stable substrates of almost any or-less cylindrical (but in some cases curved or branched) plant axes. kind appear to be the primary prerequisite for the fixation of oyster The identity of the plants to which the oysters were attached has never larvae. Some oyster communities sustain multiple generations of shell been assessed in detail. fixation to form dense reef or mound-like associations of mutually en- Here we investigate a large new collection of oyster fossils from crusting individuals (Pufahl and James, 2006). strata assigned to the informal Belemnellocamax mammillatus zone of Articulated shells of the oyster A. incurva are particularly abundant latest early Campanian age (upper third of Bed 3 in Fig. 2) at Åsen in in the upper part of Bed 3 (Fig. 2) at Åsen, and many bear indentations the Kristianstad Basin to determine whether the attached axes can be indicating attachment to cylindrical objects. The oysters included in the assigned to a particular plant group or organ. We also provide new study of Sørensen and Surlyk (2008) are all derived from this part of the palaeoenvironmental data from foraminifera and teeth, and as- sequence. These oysters form part of a relatively low diversity in- sess the composition of the fossil flora and fauna as a whole, to assess vertebrate macrofossil assemblage in Bed 3 at Åsen in contrast to the local salinity and water depth regime of shell accumulation and strikingly high diversity coeval assemblages at Ivö Klack and Ignaberga whether the interpretation of a mangrove-like depositional setting is elsewhere in the northern Kristianstad Basin. tenable for this deposit. The marine oyster-bearing deposits near the northern margin of the Kristianstad Basin are underlain by to lower Campanian floodplain to littoral deposits. These non-marine to paralic strata are 2. Geological setting best represented at Åsen, Axeltorp and Ryedal, and have yielded a broad array of charcoalified angiosperm remains (Friis, 1984, 1990; 2.1. Overview Friis and Skarby, 1981; Eklund et al., 1997; Schönenberger and Friis, 2001), together with sparse permineralized and lignitized cones The Kristianstad Basin constitutes the onshore extension of the Hanö and wood (Conwentz, 1892; Grönwall, 1915; Nykvist, 1957; Srinivasan Bay Basin (Kumpas, 1980). It is divided into two sedimentary sub-ba- and Friis, 1989), and diverse spore-pollen assemblages (Skarby, 1968; sins, demarcated to the southwest by the faulted margins of the Näv- Skarby et al., 2004). Much of this palaeoflora remains unpublished but lingeåsen and Linderödsåsen horsts (Bergström and Sundquist, 1978; taxa affiliated with modern mangroves have not been identified. Kumpas, 1980, fig. 1). Sedimentation commenced in the Overlying Campanian strata from the same basin host rich marine (Norling and Skoglund, 1977; Bergström and Sundquist, 1978; Kumpas, invertebrate and vertebrate faunas and sparse plant fossils (Erlström 1980; Norling, 1981) and borehole data indicate that subsidence and Gabrielson, 1992; Sørensen and Surlyk, 2008, 2011; Surlyk and peaked in the early Campanian (Bergström and Sundquist, 1978, fig. Sørensen, 2010; Gale and Sørensen, 2015a, 2015b; Einarsson et al., 28). Based on unpublished data, Erlström and Gabrielson (1992)

100 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

the Kristianstad Basin but a few natural outcrops of uppermost lower Glacial deposits and lower upper Campanian biocalcarenites are present along the shores of lake Ivösjön (Lundegren, 1934). Apart from natural outcrops,

QUATERN. Upper Cretaceous strata are exposed in active and disused quarries along the margins of the basin where the cover is relatively thin. The currently accessible strata range in age from early or early middle Santonian (Gonioteuthis westfalica westfalica Zone, sensu Christensen, 1997) to latest Campanian ( lanceolata Zone; see Thibault et al., 2012 and Voigt et al., 2012 for a discussion of the – 4 Campanian Maastrichtian boundary). These deposits accumulated at palaeolatitudes of 47–49°N (Kent and Irving, 2010; van Hinsbergen et al., 2015).

2.2. The Åsen area lowermost upper Previous quarrying activities at Åsen have exposed approximately 20 m of upper Santonian–lower Campanian weakly consolidated floodplain sediments rich in plant remains (see Introduction). These basal non-marine Cretaceous sediments were deposited within a NNE- trending palaeovalley, open to the south and demarcated on either side by Proterozoic igneous rocks (Lundegren, 1934; Siversson et al., 2016; see fig. 3b of Surlyk and Sørensen, 2010). The sequence rests on re- sidual kaolin clay and is overlain in places by up to 5 m of marine sand of latest early to earliest late Campanian age (Iqbal, 2013; Einarsson 3 1 Bed number et al., 2016; Siversson et al., 2016). The stratigraphy described below is CAMPANIAN based on a 3.3-m-thick section recorded by MS and the late Walter Belemnoid rostra Kegel Christensen in 2000. The uppermost lower Campanian part of the Mostly articulated marine deposit is about 2 m thick and comprises three distinct beds. The lowermost unit (Bed 1; Fig. 2) is up to 50 cm thick and consists of oyster shells poorly consolidated and very coarse grained, grey quartz Disarticulated and with charcoal and fragments of belemnites and oysters in its lower half. fragmented oyster It is truncated and overlain by 10–15 cm of green, moderately con-

uppermost lower shells solidated, coarse-grained quartz sandstone (Bed 2), rich in pre- Crystalline () dominantly marine vertebrate remains, fragmentary belemnite rostra clasts and oyster shells. Although glauconite is present throughout the Cam- 2 panian marine succession at Åsen, the green colour of Bed 2 appears to Charcoal result from impurities within the quartz grains. The stark contrast be- tween the green quartz and the white oyster shell fragments gives the 1 bed a distinctly mottled appearance, readily recognisable in the field. Bed 2 is overlain by an up to 1.5-m-thick interval of poorly consolidated

50 cm greensand, coarsening into a shell bed in the upper part (Bed 3). The shelly interval is moderately indurated and packed with mostly ar- ticulated Acutostrea shells and well-preserved belemnite rostra. The matrix in the upper, more calcareous part of Bed 3 comprises medium- SANT.- CAMP. grained quartz (sharply angular to well-rounded), inoceramid rods, Fig. 2. Stratigraphical section of the marine Campanian deposit at Åsen: (1) grey, very glauconite grains and micrite. A burrowed omission surface caps the coarse grained, quartz sand with charcoal and fragmentary belemnite rostra in the lower uppermost lower Campanian sequence. The overlying Bed 4 constitutes half; (2) green coarse-grained shelly quartz sand with fragmentary oysters (Acutostrea medium-grained, calcareous quartz sand of earliest late Campanian age incurva) and belemnites; (3) greensand grading to calcareous quartz sand packed with (Siversson et al., 2016) with a basal lag comprising reworked be- articulated Acutostrea shells and belemnite rostra in the upper third; (4) calcareous quartz sand with a basal lag consisting of reworked belemnites from Bed 3. lemnites of latest early Campanian age. The Cretaceous deposits are capped by a few metres of Quaternary moraine. indicated that the Cretaceous succession is up to 250 m thick in the 2.3. Local biostratigraphy ‘central parts of the Kristianstad Basin’. These strata persist south- eastwards into the Hanö Bay Basin beneath the where they All four marine fossiliferous units (Beds 1–4; Fig. 2) at Åsen re- reach thicknesses of around 700 m (Sopher et al., 2016). The northern cognized by Siversson et al. (2016) yield belemnite rostra but the ma- boundary of the Kristianstad Basin is erosional with numerous, isolated terial is fragmentary in Beds 1 and 2. Belemnites have proven to be outliers of Cretaceous deposits, whereas the southeastern border is ar- fundamentally important for dating the Upper Cretaceous deposits of bitrarily drawn along the coastline of the Baltic Sea (Fig. 1). The pre- the Kristianstad Basin. Christensen (1975) defined several informal, dominantly granito-gneissic basement of the northern sub-basin (the local belemnite assemblage zones, enabling correlation with the mac- Vinslöv tectonic block; see Bergström and Sundquist, 1978, fig. 31)is rofossil-based division of the Campanian in . Well-preserved capped by an undulating, deeply weathered surface dipping towards belemnite rostra occur abundantly in the upper third of Bed 3 and in- the south (Lidmar-Bergström, 1982). Weathering-resistant inselbergs of clude Belemnellocamax mammillatus (Nilsson, 1827), mu- basement rocks rise above the Kristianstad Plain in the northeastern cronata (Link, 1807) and Gonioteuthis quadrata scaniensis Christensen, part of the basin (Lundegren, 1934). Up to about 30 m of Quaternary, 1975. This assemblage is diagnostic of Christensen's (1975) informal B. glaciofluvial sediments directly overlie Upper Cretaceous strata across mammillatus zone, which correlates with the uppermost lower

101 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Campanian Belemnitella mucronata senior-Gonioteuthis quadrata gracilis seven scars per axis width (roughly half the circumference of the ori- Zone (sensu germanico). Siversson et al. (2016) presented unpublished ginal axis: Fig. 3C). The scars are up to 2 mm wide and 1 mm long (in notes by Walter Kegel Christensen of a sample of 15 rostra of B. mam- the orientation of the axis). A deeper indentation (to 0.5 mm) in the millatus from the upper third of Bed 3. The relatively deep pseu- encrusting shell has been generated on the distal side of each scar doalveolus and strongly allometric relationship between the length of (forming a blunt knob on the cast); the proximal side of each scar im- the rostrum and the depth of the pseudoalveolus in these specimens pression is less indented (Fig. 3E). A weak circular to oval tu- indicate a young age within the B. mammillatus zone (Walter Kegel bercle, < 0.5 mm in maximum diameter, is preserved near the centre of Christensen pers. com. to MS, 2000). the each scar impression (represented by a slight indentation on casts). The scars are separated by featureless rounded ridges up to 0.5 mm 3. Materials and methods wide, forming depressions on the cast.

Excavations at Åsen were carried out over many field seasons since 4.1.2. Interpretation the early 1990s, with intense systematic sampling of fossil faunas un- This mould in the oyster valve has features consistent with a conifer dertaken by a team of volunteers supervised by EE between 2010 and axis from which tightly spaced and helically arranged leaves have de- 2012. The studied section was exposed using a mechanical excavator, hisced. Conifer stems/branches with equivalent spirally arranged leaf after which bulk sampling of beds was undertaken using hand tools for scars and an associated ovuliferous cone were illustrated from coeval or wet-sieving with a 2 mm mesh. Samples were dried and handpicked for slightly older (Santonian?) fluvial from the northern margin invertebrates on site, or in the laboratory. Bulk samples of oyster shells of the Kristianstad Basin (at Ryedal) by Conwentz (1892) and assigned were sampled from Bed 3 as part of the PhD field program of EE with to a new species (Pinus nathorstii); that axis is re-illustrated here for details provided in her dissertation at the Department of Geology, Lund comparison (Fig. 3G). If affiliated with the axis, the associated ovuli- University. Oysters with enclosed sediment sampled for foraminiferal ferous cone with tightly spaced scales bearing multiple-seeds suggests investigations and specimen NRM Mo183205, bearing the impression of assignment to Cupressaceae, which was a widespread family of conifers a conifer branch, were collected by MS. Fossils illustrated in this study in the Late Cretaceous of Eurasia (Vakhrameev, 1991). are registered in the collections of the Swedish Museum of Natural History, Stockholm (NRM). Macrofossils were photographed using low- 4.2. Other oyster-encrusted plant axes angle illumination from the upper left using a Canon EOS 40D digital camera with a polarizing filter. Friable, quartz-dominated sand taken Numerous additional oyster shells were recovered during the course directly from an oyster cavity was scattered onto a sorting tray and of systematic sampling of macrofossils from the Belemnellocamax randomly selected grid squares were imaged using a biological micro- mammillatus zone at Åsen. Many show signs of unambiguous attach- scope and NIH-Elements image-capturing software. The sediment ment to plant axes, whereas others possess morphological features ty- composition was based on counting and broadly identifying all grains pical of attachment to rocks or shells. Many oyster shells are frag- within successive grid squares up to a total of 375 grains. Grain sizes mentary or the attachment features are ambiguous, hence their were measured using image analysis tools. Foraminifera were system- substrate cannot be determined with confidence. atically picked from the strewn sediment using a damp, very fine sable- Features typical of attachment to plants include: (1), encrustations hair brush. Two hundred and ninety-three specimens, picked from of roughly cylindrical form (Fig. 3F); (2), irregularities along the length successive grid squares, were sorted into species groups on a microfossil of the encrustation related to branching, leaf scars, and arching or assemblage slide. Selected uncoated specimens were imaged using a twisting of the axis (Fig. 3I–M); and (3), longitudinal surface striae Hitachi TM 3030 Plus Tabletop Scanning Electron Microscope. The il- corresponding to wood grain or bark texture (Fig. 3N). Features typical lustrated vertebrate material was collected between 1991 and 2003 by of attachment to other invertebrate hard parts include: (1), retention of Johan Lindgren (sturgeon dermal scute from Bed 2) and Mikael the original host shell in attachment to the oyster; (2), preservation of Siversson (shark teeth from the upper third of Bed 3). These specimens diagnostic shell ornament and/or growth lines in the encrustation were retrieved from the weakly consolidated sand by means of wet mould; and (3), preservation of distinct shapes denoting shell or other sieving (mesh aperture 2 mm). Because of the enormous number of invertebrate hard parts (including straight smooth cylindrical be- vertebrate specimens collected from the Åsen locality (several tens of lemnite rostra). Features typical of attachment to rocks include: (1), thousands), the collecting datum was typically not recorded for in- retention of the original encrusted pebble; and (2), irregular or crys- dividual specimens. The illustrated vertebrate fossils were coated with talline imprints on the encrustation surface. ammonium chloride sublimate prior to being photographed. Of those oysters retaining unambiguous moulds of plant axes, most can not be assigned to any specific plant group. Examples with trans- 4. Oyster–plant associations verse (Fig. 3I, K) or longitudinal striae (Fig. 3N) appear to represent wood or bark grain and are non-diagnostic for any particular woody 4.1. Oyster encrustation of a conifer branch with leaf scars plant group. Branched axes (Fig. 3K) are similarly non-diagnostic. Some moulds appear to incorporate short lateral appendages (Fig. 3L) or ir- 4.1.1. Description regular surficial scars (Fig. 3M) reminiscent of scale-leafed conifer Many oyster valves bearing scars of attachment to plant remains are shoots. Halamski et al. (2016) illustrated numerous scale-leafed conifer preserved in the upper part of Bed 3 at Åsen and some have similarities twigs (Pagiophyllum and Cyparissidium species) from other Campanian to Late Cretaceous plant axes preserved elsewhere in the Kristianstad deposits in Skåne that are similar in size and morphology to these Basin (Fig. 3A–H; Conwentz, 1892; Halamski et al., 2016). One large moulds. oyster fossil (Acutostrea incurva; NRM Mo183205) from the Be- lemnellocamax mammillatus zone at Åsen bears a 5-mm-deep imprint of 4.3. Survey of oyster attachments at Åsen a cylindrical plant axis on the underside of the left (lowermost) valve (Fig. 3A) that is so pronounced as to form a raised ridge on the opposing A survey of 2640 oyster shells > 1 cm in diameter from this deposit right valve (Fig. 3B, D, H). The axis impression is aligned more or less revealed 279 (10.6%) retained unambiguous evidence of attachment to centrally along the long axis of the oyster shell. The original axis ex- plant axes. The remainder showed attachments to other shells or in- ceeded 42 mm long and was at least 21 mm wide, retaining this width vertebrate hard parts (11.1%), rocks/pebbles (8.6%) or had ambiguous throughout the preserved specimen. The axis is covered by transversely attachment features 69.7% (Fig. 4). Of those attached to plant mate- elliptical to diamond-shaped scars arranged in tight spirals of around rials, the host axis breadth was in the range of 2–22 mm. A histogram

102 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Fig. 3. Oysters from the latest early Campanian Belemnellocamax mammillatus zone at Åsen with encrustation moulds of woody axes; and a similar sedimentary axis mould from?Santonian strata at Ryedal, northern Kristianstad Basin. A, NRM Mo183205, left valve of oyster with large axis indentation bearing numerous leaf scars. B, right valve of the NRM Mo183205 showing ridge on the right valve corresponding to the axis imprint on the left valve. C, dental polymer cast of NRM Mo183205 showing spirally arranged leaf scars. D, distal view of NRM Mo183205 showing the depth of axis indentation into the oyster valves. E, enlargement of the cast of NRM Mo183205 showing details of the leaf scars. F, oyster left valve with slightly curved cylindrical plant axis mould extending the length of the shell, NRM Mo183206. G, NRM S085156, Pinites nathorstii Conwentz, 1892 axis impression from Ryedal with equivalent tight helical rhombic leaf scar arrangement as represented on NRM Mo183205. H, lateral view of NRM Mo183205 (left valve lowermost). I, left valve with slightly sinuous axis indentation with transverse striae, NRM Mo183207. J, left valve with branched striate axis mould, NRM Mo183208. K, left valve with off-centred axis mould with transversely striate ornament, NRM Mo183209. L, portion of left valve with axis mould bearing short leaf-like oblique appendages, NRM Mo183210. M, left valve with large axis mould bearing sparse irregular leaf scars, NRM Mo183211. N, left valve with large, longitudinally striate, axis mould, NRM Mo183212. Scale bars = 10 mm. plotting the widths of attached axes (Fig. 5) has a more-or-less normal to, and centred over, the encrusted plant axis (Fig. 3I–N) suggesting size distribution suggesting moderate sorting of transported woody preferential orientation of the oyster larva to the substrate at the time of elements in the marine environment. Axes of 6–12 mm width are settling. clearly the predominant size class attached to oysters. The most dis- tinctive illustrated axis mould with diamond-shaped leaf scars com- parable to Conwentz’ (1892) Pinus nathorstii specimen (Fig. 3A) is at the maximum size limit identified for encrusted woody axes. In the vast majority of cases, the long axis of the attached oyster is roughly parallel

103 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Oysters attached Table 1 Details of sediment and contained foraminiferal assemblage from cavity of oyster NRM to shells: 11.1% Pr7786. (n = number of specimens counted; s = standard deviation; ∝ = Fisher diversity (n = 292) index) Oysters attached Component Value to rocks: 8.6% (n = 228) Sediment composition (%; n = 375) Quartz 97 Inoceramid prisms 2 Foraminifera < 1 Ostracoda < 1 Oysters Glauconite < 1 attached Indeterminate skeletal fragments < 1 to plant Quartz grain size (n = 200) Average grain size (mm) 0.247 axes: (s = 0.057) Ambiguous attachment 10.6% Foraminifera orders (% abundance; n = 293) scars 67.7% (n = 1841) (n = 279) Lagenida (benthic) 5 Buliminida (benthic 8 (benthic) 83 Globigerinida (planktonic) 5

Benthic foraminifera species (% abundance in benthic assemblage; n = 279) Lagenida Globulina lacrima (Reuss, 1845)[Polymorphina <1 (Globulina)] Fig. 4. Pie chart of the results of a survey of 2640 oyster shells > 1 cm in diameter from Guttulina trigonula (Reuss, 1845)[Polymorphina <1 the latest early Campanian Belemnellocamax mammillatus zone at Åsen showing the re- (Guttulina)] presentation of substrate attachment types (shells, rocks, woody axes and indeterminate Hemirobulina troedssoni (Brotzen, 1936)[Marginulina]<1 materials). Laevidentalina subtilis (Brotzen, 1936)[Marginulina]<1 Lagena sp. cf. L. acuticosta Reuss, 1862;? = sulcata <1 (Walker & Jacob) of Erlström and Gabrielson, 1986, fig. 100 12C Pseudopolymorphina mendezensis (White, 1928) <1 [Polymorphina (Polymorphina)] n = 279 Pyramidulina spp. ex. gr. P. obscura (Reuss, 1845) <1 mean = 7.3 mm [Nodosaria] Pyrulina sp. 1 Undifferentiated uniserial nodosariids 3 Undifferentiated polymorphines < 1 50 Buliminida Pyramidina sp. < 1 Sitella laevis (Beissel, 1891)[Bulimina]; =Praebulimina 8 reussi (Morrow) of Erlström and Gabrielson, 1986, fig. 12D =P. reussi of Gabrielson, 1991, fig. 10 Indeterminate bolivinid < 1 Rotaliida Cibicides excavata Brotzen, 1936 6 0 Dyocibicides ribbingi (Brotzen, 1936)[]; =Cibicides 2 ribbingi of Erlström and Gabrielson, 1986, fig. 12F = C. 0246810121416182022 ribbingi of Gabrielson, 1991, fig. 12

Number of specimens attached to stems Width of attached axis (mm) Eoeponidella sp. < 1 Gyroidinoides umbilicatus (d'Orbigny, 1840) 10 Fig. 5. Histogram of the widths of woody axes based on 279 oysters from the latest early [Rotalina] = G. umbilicatus of Gabrielson, 1991, fig. 14 Campanian Belemnellocamax mammillatus zone at Åsen with unambiguous attachments to Lingulogavelinella sandidgei (Brotzen, 1936)[Cibicides]; 54 fi plant substrates. =Gavelinella spp. of Erlström and Gabrielson, 1986, g. 12A, B = G. sandidgei of Gabrielson, 1991, fig. 13 Nonionella warburgi Brotzen, 1936 5 5. Foraminiferal evidence of palaeoenvironment Indeterminate rotaliids 10 Morphotypes present among Rotaliida (% abundance; n = 242) 5.1. Epibiont Foraminifera on oysters Planoconvex morphotypes 96 Biconvex morphotypes 4

Foraminifera belonging to Dyocibicides ribbingi (Table 1; Figs. 6A, B; Planktonic foraminiferal species (total number specimens = 14) 7R) are firmly attached to the surfaces of a few specimens of Acutostrea Archaeoglobigerina blowi Pessagno, 1967 incurva. Clusters of D. ribbingi, each with a few specimens, are common Costellagerina pilula (Belford, 1960)[Rugoglobigerina] Globigerinelloides prairiehillensis Pessagno, 1967;=G. multispina on the ventral inner surface of one of the oysters (NRM Mo183201; (Lalicker) of Erlström and Gabrielson, 1986, fig. 12E Fig. 6A, B). The pattern of encrustation suggests that the foraminiferans ? sp. are in their original growth positions. On other oysters (e.g., NRM globulosa (Ehrenberg, 1839)[Textularia] Mo183203, NRM Mo183204), isolated specimens of D. ribbingi are ce- Foraminiferal maximum dimensions (average and standard deviation; mm) mented on external surfaces. No other epibonts were identified on oy- Lagenida (benthic) [n = 15] 0.44 (s = 0.20) sters examined for Foraminifera. The surfaces of the oysters also show Buliminida (benthic) [n = 22] 0.21 (s = 0.03) no evidence of minute borings for cyanobacterial attachment or larger Rotaliida (benthic) [n = 242] 0.23 (s = 0.05) Globigerinida (planktonic) [n = 14] 0.26 (s = 0.07) cavities caused by the holdfasts of sessile soft-bodied marine organisms or by other borers. Benthic foraminiferal diversity (Fisher diversity index calculated from Murray, 1991, fig. 2.2) Brotzen (1936) originally described Dyocibicides ribbingi from the (continued on next page)

104 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Table 1 (continued) assemblage, forming 54% of benthic specimens. This species was ori- ginally described by Brotzen (1936) from the Upper Cretaceous of Component Value Eriksdal and referred by him to Cibicides. Others, including Erlström Estimated ∝ diversity (based on identified species + estimate 4–6 and Gabrielson (1986) and Gabrielson (1991), transferred the species to of indeterminate species) Gavelinella Brotzen, 1942 ( species, Discorbina pertusa Marsson, 1878; see lectotype figured by Revets, 1996, pl. 7, figs. 1–4). Brotzen's species lacks the deep umbilicus of G. pertusa and has more extensive Upper Cretaceous at Eriksdal, Sweden. He referred the species to Cibi- umbilical flaps that, except where broken (as in Fig. 7ZA), cover the cides de Montfort, 1808 (type species, Cibicides refulgens de Montfort, umbilical region (Fig. 7Z, ZD). The umbilical flaps also distinguish the ff 1808). In terms of generic characters, it di ers from C. refulgens in its species from Cibicides. In peripheral view, the profile of the final restricted low interiomarginal arched aperture on the dorsal (umbilical) chamber in L. sandidgei (with peripheral angle directed towards the side. The aperture does not extend across the periphery and along the ventral side) differs from that in G. pertusa (with peripheral angle di- spiral suture on the attached ventral side as it does in C. refulgens. Rare rected towards the dorsal side). Brotzen's species more closely re- specimens (Fig. 6A, B) associated with typical morphotypes on oyster sembles Lingulogavelinella albiensis Malapris, 1965, the type species of NRM Mo183201 develop a loose, biserial arrangement of adult cham- Lingulogavelinella Malapris, 1965 (see specimen from type area figured fi bers. Because of this, and the apertural position and evidence for rm by Revets, 1996, pl. 9, figs. 5–8), than the type species of Cibicides or cement attachment, the species is referred to Dyocibicides Cushman and Gavelinella. Valentine, 1930 (type species, Dyocibicides biserialis Cushman and Other species with abundances of over 5% in the benthic assem- Valentine, 1930). blage include Gyroidinoides umbilicatus (10%), Sitella laevis (8%), Cibicides excavata (6%) and Nonionella warburgi (5%). Rare forms be- 5.2. Foraminiferal assemblage recovered from sediment in the oyster body long within Dyocibicides, Eoeponidella, Globulina, Guttulina, cavity Hemirobulina, Laevidentalina, Lagena, Pseudopolymorphina, Pyramidina, Pyramidulina and Pyrulina (Table 1). Well-sorted, medium-grained quartz sand filling the cavity of oyster Planktonic species comprise 5% of the foraminiferal assemblage and NRM Mo183202 contains well-preserved foraminiferans that belong include species of Archaeoglobigerina, Costellagerina, Globigerinelloides to > 17 benthic species and at least 5 planktonic taxa (Table 1; and Heterohelix, together with types tentatively referred to Hedbergella. Figs. 7A–AF, 8A–L). The sand is friable and incipient cement is present The assemblage lacks zonal indices but is consistent with an early only in some areas near the oyster shell. Foraminifera form < 1% of the Campanian age. sediment. The specimens show negligible signs of dissolution or abra- Gabrielson (1991, p. 15, 33) provided a census of Foraminifera from ‘ ’ sion and are not filled by cement or sediment. Although the average test marine sand rich in plant debris and glauconite in the Åsen section. size for all groups except the lagenids is small and approximates the Although a species list was not provided and the Foraminifera were average sediment grain size, suggesting sorting, the presence of much assigned only to broad groups, it is apparent that his assemblage is si- fi larger ostracods and inoceramid prisms and the lack of abrasion or milar in composition to that recorded here from the sediment lling the fracturing of the delicate tests suggest that the microfossils have not oyster. undergone extensive transport from their life environment. The benthic assemblage has low diversity (Fisher α diversity index 6. The chondrichthyan faunas of 4–6; see Murray, 2006, p. 18, 19). During the Late Cretaceous, for- aminiferal assemblages acquired a modern aspect in terms of the orders Isolated remains of chondrichthyans (primarily shed teeth but also represented (Loeblich and Tappan, 1987, 1992). Many modern genera vertebrae, pieces of prismatic cartilage, dermal thorns and fin spines) also appeared during this time. Only three orders are present in the occur abundantly throughout Beds 1–3 but are particularly common in studied sample: the Lagenida with more than eight species form 5% of the upper half of Bed 1, Bed 2 and in the upper, highly calcareous third the total assemblage; the Buliminida with three species form about 8%; of Bed 3 (see Siverson, 1993 and Siversson et al., 2016 for a list of about and the Rotaliida with at least six species form 83%. Most of the ro- 38 species). Tooth plates and fin spines of chimaeroids are common taliids are planoconvex morphotypes that probably lived as mobile finds in Beds 1–3, and four species have been identified (Evgeny Popov, epibionts on firm substrates. Lingulogavelinella sandidgei dominates the pers. com.). Two of these occur in the upper part of Bed 3;

Fig. 6. Epibiont foraminiferans on the inner surface (distal from hinge) of oyster NRM Pr7785. The foraminiferans are variants of Dyocibicides ribbingi (Brotzen, 1936) and in growth positions. The quartz grains and inoceramid prisms that partly obscure the tests are attached by incipient diagenetic cement. Note the smooth surface of the interior of the oyster and lack of cyanobacterial attachment borings. Bar scales = 1 mm.

105 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Fig. 7. Benthic foraminiferal assemblage from sand filling the internal cavity of oysters NRM Pr7785 (Figs. F, K, L, O, S–U, Z) and NRM Pr7786 (Fig. 2A–E, G–J, M, N, P–R, W–Y, ZA–ZF). Bar scales = 0.1 mm. A, Laevidentalina subtilis (Brotzen, 1936), NRM Pr7787; B, Marginulina troedssoni Brotzen, 1936, NRM Pr7788; C, D, Pyramidulina spp. ex. gr. P. obscura (Reuss, 1845), C NRM Pr7789, D NRM Pr7790 6; E, Globulina lacrima (Reuss, 1845), NRM Pr7791; F, Guttulina trigonula (Reuss, 1845), NRM Pr7792; G, Pseudopolymorphina mendezensis (White, 1928), NRM Pr7793; H–J, Pyrulina sp., H NRM Pr7794, I NRM Pr7795, J NRM Pr7796; K, Lagena sp. cf. L. acuticosta Reuss, 1862, NRM Pr7797; L–N, Sitella laevis (Beissel, 1891), L NRM Pr7798, M NRM Pr7799, N NRM Pr7800; O, Pyramidina sp., NRM Pr7801; P, Q, Cibicides excavata Brotzen, 1936, P NRM Pr7802, Q NRM Pr7803; R, Dyocibicides ribbingi (Brotzen, 1936), NRM Pr7804; S, T, Nonionella warburgi Brotzen, 1936, S NRM Pr7805, T X NRM Pr7806; U–W, Eoeponidella sp., U NRM Pr7807, V and W NRM Pr7808; X–ZD, Lingulogavelinella sandidgei (Brotzen, 1936), X and Y NRM Pr7809, Z, NRM Pr7810, ZA NRM Pr7811, ZB and ZC NRM Pr7812, ZD NRM Pr7813; ZE, ZF, Gyroidinoides umbilicatus (d'Orbigny, 1840), NRM Pr7814. bifurcatus Case, 1978 and Edaphodon sp. Most elasmobranch species Siverson, 1992 and “” tenuis (Davis, 1890) above Bed 2, and occur in all three beds but some have a more restricted stratigraphic the vast majority of the teeth of these species are derived from the distribution. There are no definite records of hybodonts (see Rees, upper half of Bed 1. Sieving of the upper part of Bed 3 (2 mm aperture 1999), Scapanorhynchus perssoni Siverson, 1992, Anomotodon hermani mesh) yielded a highly diverse chondrichthyan assemblage,

106 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Fig. 8. Planktonic foraminiferal assemblage from sand filling the internal cavity of oysters NRM Pr7785 (Figs. A–C, E, F, I–L) and NRM Pr7786 (Figs. D, G, H). Bar scales = 0.1 mm. A, Heterohelix globulosa (Ehrenberg, 1839), NRM Pr7815; B–D, Globigerinelloides prairiehillensis Pessagno, 1967, B and C NRM Pr7816, D NRM Pr7817; E, F Hedbergella? sp., NRM Pr7818; G, H, Archaeoglobigerina blowi Pessagno, 1967, NRM Pr7819; I–L, Costellagerina pilula (Belford, 1960), I NRM Pr7820, J NRM Pr7821, K NRM Pr7822, L NRM Pr7823. numerically dominated by teeth of two benthic taxa; an undescribed belemnites, represented mainly by rostra of B. mammillatus species of Heterodontus (molariform teeth of H. portusjacksoni-type) and (Christensen, 1975; Sørensen and Surlyk, 2008; Einarsson pers. obs.). Squatina lundegreni Siversson et al., 2016. Additional, common benthic Other molluscs in Bed 3 at Åsen include small inoceramid bivalve species include an undescribed, large-toothed Squatirhina Casier, 1947 fragments (Lundgren, 1876; Sørensen and Surlyk, 2008; Einarsson pers. and casieri Herman in Cappetta and Case, 1975. The lam- obs.), Trigonia buchi Geinitz, 1840 (=Apiotrigonia sulcataria de Lamarck, niform shark “Carcharias” latus (Davis, 1890) is the third most common 1819—a colonist of coarse-grained substrates: Carlsson, 1938), and elasmobranch species and is represented by a few large teeth and a free-living pectinids. Microscopic inoceramid prisms constitute a sig- much larger number of teeth from small, presumably juvenile, in- nificant component of the sediment's sand fraction (Siversson et al., dividuals. Teeth of large (> 2 m estimated body length), apex-predator 2016; Fig. 6A, B). There is no evidence of infaunal mollusc species or sharks are common and represented primarily by Archaeolamna ko- gastropods in Bed 3 at Åsen (Moberg, 1884; Lundgren, 1894; pingensis (Davis, 1890), Cretalamna sarcoportheta Siversson et al., 2015 Lundegren, 1934; Einarsson pers. obs.), although this may be a con- and lindstromi (Davis, 1890). Teeth of A. kopingensis and S. sequence of differential preservation (e.g., preferential dissolution of lindstromi outnumber those of C. sarcoportheta by a large margin in Beds aragonitic shells or a sediment type not conducive to the preservation of 1 and 2, whereas teeth of the latter are about equally abundant as those body chamber casts). are sparse and represented only by of the other two combined in the uppermost, Acutostrea-rich part of Bed cranolaris (Sørensen and Surlyk, 2008) and the ichnogenus 3. (pedicle boring) Podichnus (Sørensen and Surlyk, 2008; Einarsson pers. obs.) in contrast to nearly 30 species encrusted on and below the 7. Other invertebrate fossils in Bed 3 boulders of the coeval rocky coastal deposit at Ivö Klack (Christensen, 1969; Surlyk and Sørensen, 2010). Scleractinian are represented The invertebrate macrofossil assemblage in Bed 3 is strongly in Bed 3 at Åsen by the free solitary ahermatypic suspension-feeding dominated by articulated and disarticulated valves of A. incurva, and Micrabacia hilgardi (Stephenson, 1916). In contrast, 11 scleractinian

107 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114 species are represented in the rocky shoreline deposits at Ivö Klack, one settings of the Northern Hemisphere (Sender et al., 2015), and has been of which is the congeneric Micrabacia suecia (Surlyk and Sørensen, interpreted to have attained a mangrove-like habit at some localities 2010). are represented only by phosphatic fragments of (Schweitzer et al., 2003; Abu Hamad et al., 2016; El-Saadawi et al., propodi from indeterminable decapods (Einarsson pers. obs.) and by 2016). However, others have interpreted this plant to characterize su- plates (Arcoscalpellum)(Sørensen and Surlyk, 2008). At Ivö pratidal coastal marsh communities (Silantieva and Krassilov, 2006), Klack, decapods are represented by Protocallianassa faujasi (Einarsson and it appears to have been rare or absent after the Cenomanian. The et al., 2016) and are represented by 28 species encrusted on herbaceous mangrove , Acrostichum, is known from the Maas- shells, pebbles and boulders (Gale and Sørensen, 2015a, 2015b). trichtian but has been recorded mainly from Cenozoic strata (Bonde Echinoids are represented by fragmentary plates and spines, one as- and Kumaran, 2002). The rhizomes of Acrostichum differ from the Åsen signed to a cidaroid echinoid (Sørensen and Surlyk, 2008; Einarsson axes in being clad in robust scales (Bonde and Kumaran, 2002). pers. obs.). In contrast, echinoids are represented by the fragmented Seagrasses also have a poorly constrained evolutionary history with remains of 18 species at Ivö Klack (Surlyk and Sørensen, 2010). A few earliest records again from the Maastrichtian (van der Ham et al., bryozoans and serpulid worm tubes have been reported to encrust the 2017), but this group is unlikely to have contributed the plant material surface of oyster and belemnite rostra in Bed 3 at Åsen (Sørensen and on which the Åsen oysters were encrusted because they do not produce Surlyk, 2008). At Ivö Klack, are represented by 17 species robust woody tissues. encrusting both boulders and bivalves (Surlyk and Sørensen, 2010). The collective palaeobotanical evidence suggests that the main ra- Overall, although oyster shells, belemnite rostra and inoceramid frag- diations of both modern mangrove and seagrass ecosystems were pre- ments are abundant, the invertebrate fauna at Åsen is of very low di- dominantly Paleogene phenomena (Plaziat et al., 2001). Although the versity compared with coeval deposits at Ivö Klack. fossil pollen record of angiosperms extends back to the in Sweden (Vajda, 2001; Vajda and Wigforss-Lange, 2006), there are no 8. Discussion records of Cretaceous mangrove pollen from this region. There is no strong palaeobotanical evidence for the development of a mangrove 8.1. Were there mangroves in the Campanian of the Kristianstad Basin? habit among angiosperms or conifers in Late Cretaceous ecosystems globally prior to the Maastrichtian. Although a few plant groups arguably developed a mangrove-like habit as early as the late Palaeozoic, e.g., Cordaitales (Raymond and 8.2. Plants as a substrate for oysters Phillips, 1983) and glossopterids (Jafar and Kar, 1996), these plants grew predominantly in freshwater settings or were only episodically Based on the quantitative survey of oyster shells collected from the exposed to marine influence (McLoughlin, 1993, 2011; Falcon-Lang, Belemnellocamax mammillatus zone, small woody plant axis fragments 2005) and lack the specialized adaptations to salinity characteristic of were common components of the bottom-water communities at Åsen modern mangroves (Tomlinson, 1986). A few plant groups and provided an important substrate for encrusting oysters. The small have also been interpreted to have developed mangrove-like lifestyles, axes with a more-or-less normal size distribution, together with the e.g., Early Triassic pleuromeian lycopsids (Krassilov and Zakharov, presence of unambigiuous leaf scars on some specimens, indicates that 1975; Retallack, 1975, 1977), some cheirolepidiacean conifers the woody materials mostly, or perhaps entirely, represent distal por- (Philippe, 1992), the Cenomanian ginkgoalean Nehvizdyella (Falcon- tions of subaerial axes that were rafted into the marine environment. It Lang, 2004; Kvaček et al., 2005), and the Maastrichtian putative her- is unlikely that they represent mangrove pneumatophores, which tend baceous lycopsid Mosacaulis (van der Ham et al., 2011). However, these to be larger on average, have a narrow size range, and incorporate a groups have quite distinctive stem morphologies or lack the robust spongy, lenticel-rich, outer cortex and/or aerenchyma (Kathiresan and longitudinally striate axis ornamentation or tight rhombic leaf scar Bingham, 2001), or mangrove prop roots or fasciculate roots, which helices characteristic of the stems encrusted by oysters at Åsen. More- tend to have a broader size range than the available material and are over, the evidence supporting a mangrove habit for most of these typically represented by discrete size classes associated with their groups is equivocal and is generally based on the simple preservation of modular architecture of successive branching (Farnsworth and Ellison, dispersed plant remains in invertebrate-rich marine sedimentary facies 1996). (Plaziat et al., 2001). We envisage that the small plant remains encrusted by oysters in the Modern angiospermous mangroves do not constitute a mono- uppermost lower Campanian Belemnellocamax mammillatus zone at phyletic group. Rather, the mangrove habit, involving specialized Åsen in the Kristianstad Basin were predominantly decorticated sub- morphological adaptations to salt tolerance, low-oxygen substrates, and aerial woody plant axes, or in a few cases, retained bark and leaf scar marine transportation of propagules, developed independently in arrangements. Most examples appear to represent fragments of small around 20 families (Tomlinson, 1986; Ellison et al., 1999). Ellison et al. distal woody branches that were washed into an open subtropical (1999) summarized the fossil record of mangroves and noted that the marine embayment. Abundant charcoal in Bed 3 (Einarsson pers. obs.) earliest confident examples of modern mangrove-forming families are also attests to a transported origin of the woody material at Åsen. The from the Maastrichtian. Nypa (Arecaceae), the mangrove palm, has the relative scarcity of terrigenous pebbles at the depositional site (sug- longest stratigraphic range among modern mangroves (Maastrichtian or gesting a location of low energy and/or moderate distance from the possibly latest Campanian to present) based on fossil pollen, fruits and shore) accounts for the importance of detrital shell and woody mate- leaves (Schrank, 1987; Thanikaimoni, 1987; Collinson et al., 1993; rials as colonization substrates for oysters in the local environment. The Ellison et al., 1999; Gee, 2001; El-Soughier et al., 2011) but this genus discrete size range of encrusted woody plant fragments suggests that has very straight roots and stems lacking the small helical leaf scars either (1), the oyster juveniles had a preference for encrusting woody evident on the largest of the Åsen specimens (Fig. 3A, C). Palaeo- substrates of this specific diameter, or (2), these were the sizes of plant wetherellia (Euphorbiaceae) also has a record extending back to the remains most commonly represented in the oyster-rich environment Maastrichtian in (Mazer and Tiffney, 1982), probably re- owing to modest current-winnowing and sorting of sedimentary mate- presenting the earliest dicotyledonous angiosperm mangrove. Other rials. families developed representatives with a mangrove habit at various times through the Cenozoic (Ellison et al., 1999). 8.3. Environment suggested by Foraminifera The short matoniacean tree-fern Weichselia reticulata (Stokes et Webb) Fontaine in Ward emend Alvin, 1971 has been documented from The foraminiferal epibionts (Section 5.1) colonized the surface of to Cenomanian lowland freshwater swamp to coastal the oysters after death of the bivalves. The benthic assemblage in the

108 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114 sediment filling the oyster cavity has not been transported far from the Citharina suturalis, Anomalinoides harperi and Nonionella sp. None of life environment (Section 5.2), but the presence of planktonic tests these species has been found at Åsen. Deeper water biofacies in the suggests incorporation of some material from outside the immediate palaeoslope model include carbonate-cemented agglutinated species oyster community. The foraminiferal assemblage recorded previously together with a diverse array of calcareous hyaline taxa. Calcareous- from sandy facies in the Åsen section is similar to that recorded here cemented agglutinated Foraminifera, such as Heterostomella, the lagenid from sand within an oyster shell (Section 5.2). During encrustation by Neoflabellina, the biserial buliminid Bolivinoides, and the rotaliids the foraminiferal epibionts and infilling by sediment, the valves of the Osangularia, Stensioeina and Quadrimorphina that Gabrielson (1991), oyster remained articulated but partly open and then were tightly recorded from other localities in the Kristianstad Basin, are common in closed during subsequent burial beneath sand. The articulation of the the mid-neritic to bathyal biofacies of the palaeoslope model developed oyster shells suggests that great water-current action was not involved by Olsson and Nyong (1984). The absence of these taxa at Åsen sup- in the deposition. The foraminiferal assemblage can, therefore, be used ports the hypothesis that the depositional environment here was shal- in interpretations of the depositional environment. Environmental as- lower in the inner neritic zone. The differences in composition between sessment is made using (1) modern analogues; and (2) Campanian the North American and Åsen inner neritic assemblages are probably foraminiferal distributions along a palaeoslope model and distributions the result of salinity differences between the restricted Åsen embay- backed by facies analyses. ment and the open-marine New Jersey-Delaware shelf. Foraminiferal distributions in modern mangrove environments and The Foraminifera at Åsen, therefore, suggest an offshore inner on the adjacent marine shelf have been studied extensively (e.g., Todd neritic environment with a clean sand substrate. It is significant that no and Brönnimann, 1957; Murray, 1991, 2006; Berkeley et al., 2009a; small gastropods, common among mangroves, are present in the sedi- Satyanarayana et al., 2014; Eichler et al., 2015; Fajemila et al., 2015; ment filling the oyster cavity; and, apart from Dyocibicides, no other Jeyabal et al., 2016; Sen et al., 2016). The Åsen assemblage differs from epibionts (e.g., serpulids or bryozoans as figured by Erlström and modern mangrove-associated assemblages by: (1) a lack of organic- Gabrielson (1986) on a cobble found in conglomerate at Ullstorp) were cemented siliceous agglutinated Foraminifera, usually common con- observed on the oyster shells at Åsen. This supports the interpretation of stituents of modern mangrove communities; (2) having 5% Lagenida a relatively offshore rather than marginal-marine environment (the including a diversity of genera, whereas modern mangrove assemblages nearest rocky shoreline being about 1500 m to the northeast based on have very few if any Lagenida; (3) having 8% triserial Buliminida, local exposures of basement rocks). whereas similar morphotypes are usually absent from the modern mangrove assemblages but are present in deeper-water environments; 8.4. Chondrichthyans as palaeoenvironmental indicators and (4) having about 5% planktonic foraminiferans, whereas these are absent or extremely rare in modern mangrove environments. The per- Modern mangroves are utilized as nurseries by various elasmo- centage of planktonic types, their shell forms and test sizes, suggest that branchs (White and Potter, 2004) but, strictly speaking, most of these the depositional environment faced an open deeper sea, but was within species have been recorded from seagrass meadows and sand flats the inner neritic zone (based on the model of Murray, 1976). The ex- fringed by mangroves, rather than directly among the mangrove roots cellent preservation of foraminiferal tests at Åsen is also different from during high tide. Lemon sharks (Negaprion spp.) are notable exceptions, that commonly found in organic-rich mangrove settings where partial as juveniles are commonly encountered in the latter environment dissolution producing at least surficial etching of calcareous shells is (Legare et al., 2015). Some batoids and smaller sharks may remain in common (Berkeley et al., 2009b). Compared to the modern distribu- mangrove-lined bays and estuaries into the adult stage (Last and tions charted by Todd and Brönnimann (1957) in shallow-water sedi- Stevens, 2009). Predator avoidance appears to be a main factor indu- ment of the Gulf of Paria (Trinidad), the depositional water depth at cing this behaviour (Heupel and Hueter, 2002) but mortality rates from ff – Åsen best corresponds to the o shore zone (about 15 30 m) based on may nonetheless be high in the first (Manire and Gruber, the abundances of polymorphines, Nonionella, triserial Bulimina (ana- 1993; Gruber et al., 2001). Larger sharks are the main predators of logue of Sitella), and globigerinids. Planoconvex rotaliids are not as juvenile elasmobranchs (and adults of similar small size) and the latter diagnostic because they are common in a broad range of environments. are, therefore, often found in very shallow water (Springer, 1967). ff ff The di erences in benthic species diversity may be due to di erent Given that many of the Acutostrea valves in Bed 3 are preserved ar- ff substrates, with higher diversity in the muddy sediment in the o shore ticulated, it is reasonable to assume that they were exposed to minimal zone in the Gulf of Paria than in the medium-grained sand facies at transportation before burial. If the oysters grew on mangrove roots in Åsen. In the Gulf of Paria, Todd and Brönnimann, (1957, p. 21) noted the intertidal zone, the chondrichthyan remains contained within the that in all depth zones from the mangroves to about 30 m, the benthic sediment should reflect a fauna well adapted to this particular en- ‘ foraminiferans are of exceptionally small size, their maximum dimen- vironment. ’ sions rarely reaching 0.4 mm . The Åsen assemblage has similar test The chondrichthyan material collected from Bed 3 at Åsen was re- sizes (Table 1). As Todd and Brönnimann (1957) concluded for their covered by bulk sieving. Because of the relatively large aperture mesh modern assemblages, environmental factors were probably sub-optimal (2 mm), meaningful studies of population dynamics are restricted to fl for Foraminifera and may have been in uenced by lower than normal large-toothed lamniform sharks. Four of the larger lamniforms in Bed 3 fl salinity related to periodic in ux of freshwater via streams and by lower are represented by abundant material: Archaeolamna kopingensis (Davis, than normal food supply. The palaeogeography of the Kristianstad 1890), “Carcharias” latus (Davis, 1890), Cretalamna sarcoportheta Basin margin suggests the presence of a river mouth only a few kilo- Siversson et al. (2015) and Squalicorax lindstromi (Davis, 1890). Teeth — metres to the north of the Åsen locality (Siversson et al., 2016) the of A. kopingensis are common in Bed 3 but are of smaller size compared fl transgressive marine deposits at this site resting directly on uvial clays with specimens from Beds 1 and 2, indicating a larger percentage of and gravels hosting large logs (Nykvist, 1957) and leaf/charcoal lenses immature individuals (Fig. 9A). Cook et al. (2011) indicated an original (Friis and Skarby, 1981). total length of 3–4 m for a partial skeleton of this species (LACM ffi It is di cult to equate the Åsen assemblage with one of the benthic 128125) from the Campanian , western , USA. Based foraminiferal biofacies recognized in the palaeoslope model established on tooth size (using LACM 128125 as a guide) we estimate that most by Olsson and Nyong (1984) for the Campanian open-shelf succession teeth of A. kopingensis from Bed 3 are derived from individuals in the on the New Jersey–Delaware margin of North America. The shallowest 2–3 m range and represent a mix of larger juveniles and possibly also biofacies in that scheme is attributed to the inner shelf and character- adult males. The absence of teeth from very small juveniles and large ized by dominant Anomalinoides pinguis, Gavelinella nelsoni, Lenticulina adults (potentially gravid females) suggests that the area was not a pseudosecans, Globulina gibba and less common Citharina multicostata, primary nursery for this species.

109 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Fig. 9. Selected vertebrate remains from the uppermost, oyster-rich part of Bed 3 (A–G) and Bed 2 (H), B. mammillatus zone, Åsen locality. Scale bars = 10 mm (B–E, H), 3 mm (A, G), 2 mm (F). A, Archaeolamna kopingensis (Davis, 1890), upper commissural tooth from large juvenile or small mature individual, labial view, NRM P 16433, B, Cretalamna sarcoportheta Siversson et al., 2015, upper jaw tooth, possibly LP3 (see Siversson et al., 2015, Fig. 19c), labial view, NRM P 16432, C, Squalicorax lindstromi (Davis, 1890), labial view, NRM P 16431, D–E, “Carcharias” latus (Davis, 1890), D, second(?) lower anterior tooth of large individual, labial view, NRM P16429, E, first(?) upper anterior tooth of juvenile individual, labial view, NRM P16430, F, Heterodontus sp., occlusal view, NRM P16435, G, Heterodontus sp. aff. H. rugosus (Agassiz, 1839), occlusal view, NRM P 16434, H, Acipenseridae sp. (sturgeon), fragment of large dermal plate, NRM P16428.

Teeth of the otodontid shark Cretalamna sarcoportheta are abundant “Carcharias” latus is the most common lamniform species in Bed 3 (about as common as the teeth of A. kopingensis and S. lind- throughout Beds 1–3(Fig. 9D–E). The material is strongly dominated by stromi combined) and their size, relative to the maximum recorded for teeth two-thirds the tooth width or less of the largest equivalent teeth this species, indicates that the vast majority originate from large juve- (similar jaw position) of this species collected from the B. mammillatus niles to adult individuals. Assuming a low tooth count similar to that of zone of the Kristianstad Basin (Fig. 9E). The remainder of the teeth C. hattini Siversson et al., 2015, we estimate that the combined width of include the largest teeth collected of this species (Fig. 9D) and may all upper jaw teeth was about 145–150 mm on each palatoquadrate in represent large adult females (females reach a larger maximum size in large individuals (see Siversson et al., 2015, fig. 19c). Using modern all living, macrophagous lamniform sharks; see Compagno, 2001). white sharks as a template (see Leder et al., 2016, for an explanation of Modern shark nurseries are characterized by the presence of gravid the method of calculating TL based on upper jaw tooth width), this females, neonates, young-of-the-year and older juveniles (Castro, would translate into a total length closer to 2.5 m with exceptionally 1993). The size distribution of teeth of comparable jaw position in ‘C’. large individuals approaching 3 m in length. The broad, triangular, latus is, therefore, indicative of a nursery area. Young-of-the-year and upper jaw teeth (Fig. 9B) indicate a high trophic level relative to body older juveniles of the sole extant species of Carcharias, C. taurus size. Rafinesque, 1810, occur predominantly around rocky, high-profile reefs Teeth of Squalicorax lindstromi are common in Bed 3 and the ma- in shallow, coastal water where predation pressure from larger sharks is jority of them probably derive from adult individuals. The largest tooth reduced (Dicken et al., 2006). There is no evidence of Campanian rocky is 20.8 mm wide, which is the maximum tooth width recorded for the reefs of any type in the Åsen area; the nearest rocky shoreline would species in the B. mammillatus zone (Fig. 9C). Based on associated ver- have been located at least 1500 m to the northeast. Reduced visibility tebral centra from the Ignaberga locality, Bazzi et al. (2017) estimated a from seasonal riverine discharge from the north may have afforded maximum total length of at least 3.8 m for this species. Because of its some degree of protection from predators. Also, the abundance of teeth abundance, tiger shark-like dentition and large body size, S. lindstromi from juvenile ‘C’. latus and those of presumably adult, apex-predatory S. may have occupied the highest trophic level in the area during de- lindstromi and C. sarcoportheta does not necessarily mean that they all position of Bed 3. Teeth of large are relatively common in congregated in the area at the same time of the year. Bed 2 but almost all remains extracted from the upper part of Teeth of benthic elasmobranchs outnumber those of pelagic species Bed 3 originate from small individuals of propython Cope, 1869 in the upper, Acutostrea-rich part of Bed 3. The great majority of the or from juvenile to presumably adult individuals of the small-sized tooth specimens of benthic elasmobranchs belong to an undescribed Eonatator sternbergi (Wiman, 1920). species of (Heterodontus sp.; Fig. 9F) and the recently

110 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114 described angel shark Squatina lundegreni. Because of the relatively abundance of teeth of A. kopingensis up section thus indicates a trans- large aperture mesh (2 mm), the collection of Heterodontus is strongly gressive sequence. There are no modern examples of intertidal man- size biased, almost exclusively restricted to large molariform teeth of grove environments in which bullhead sharks and angel sharks dom- adult morphology (Fig. 9F–G). Small-scale, preliminary sampling using inate the elasmobranch fauna. The composition of the benthic a finer mesh indicates that juveniles were also common. Modern bull- component of the chondrichthyan fauna is, therefore, likewise at odds head sharks are inactive during the day, often seeking shelter in caves with a mangrove setting. Rather than indicating a protected mangrove in rocky reef areas (Compagno, 2001). Juveniles of H. portusjacksoni, environment, the chondrichthyan fauna suggests deeper, highly pro- one of the best studied species, stay in shallow bays and estuaries for ductive waters, possibly seasonally murky. The abundance of remains several . Adults are segregated by sex and reside in deeper offshore from bullhead sharks indicates rich feeding grounds with an assortment water (McLaughlin and O'Gower, 1971). After dusk, the sharks emerge of benthic invertebrates. Given the absence of evidence of rocky reefs in from their hiding spots in search of primarily invertebrate prey. Sea the area, it is likely that the sandy seafloor had patches of macroalgal urchins are a favourite prey for several species (see Compagno, 2001) beds, providing some cover for smaller benthic sharks. but echinoids may be missing altogether in the diet of some local po- pulations (Segura-Zarzosa et al., 1997). Smaller bony fish are also 8.5. Other invertebrates as palaeoenvironmental indicators consumed occasionally (Compagno, 2001). Unlike, for example, eagle rays, modern bullhead sharks do not prey extensively on oysters, pre- Other elements of the macroinvertebrate fauna at Åsen also favour sumably because they lack plate-like tooth batteries capable of pro- an oyster-bank depositional setting offshore from the littoral zone. cessing these thick-shelled bivalves. Although isolated plates and spines Decapods lived as burrowers within the soft substrate surrounding the of echinoids do occur in Bed 3 (Iqbal, 2013) they are probably not oyster accumulations, denoting marine conditions with water common enough to explain the great abundance of teeth from bullhead depths < 100 m (Einarsson et al., 2016). Barnacles, bryozoans, solitary sharks. In contrast to the uppermost lower Campanian sequence at corals and serpulid worms encrusted predominantly onto the oyster Åsen, echinoids are abundant at other localities in the Kristianstad shells and also onto skeletal material from pelagic belemnites and Basin where the dominant lithology is biocalcirudite and biocalcar- marine vertebrates that accumulated through fallout from the water enite. Teeth of bullhead sharks are common in the well-sampled bio- column or sediment winnowing by currents. Sørensen and Surlyk calcarenite deposits at Ignaberga (see Siverson, 1992, text-fig. 1) but (2008) identified a pedicle boring (Podichnus) in an oyster the vast majority belong to Heterodontus sp. aff. H. rugosus, a species valve from Åsen. Despite their interpretation of a mangrove-like de- that is quite rare at Åsen (Fig. 9G). The great abundance of teeth from positional setting for this oyster-rich bed, Sørensen and Surlyk (2008) bullhead sharks and other durophagous elasmobranchs in Bed 3 (e.g., noted that, apart from the infaunal Lingula, brachiopods are not typical the large shovelnose ray Rhinobatos casieri Herman in Cappetta and of modern mangal biomes. The association of relatively well-preserved Case, 1975) imply abundant invertebrate prey for which there is limited planktonic and benthic Foraminifera together with epifaunal colonial fossil evidence (e.g., , shrimp, aragonitic-shelled bivalves and organisms and terrigenous wood fragments also has similarities to some gastropods, and polychaetes). of the low-energy neritic environments of the Cenomanian–Coniacian Extant angel sharks occur on muddy and sandy bottoms, from the greensand–carbonate transition within the Fennoscandian Border Zone intertidal zone to depths up to 1300 m on the continental slopes of southern Sweden and Bornholm (Vajda-Santivanez and Solakius, (Compagno, 1984). Estuarine fronts are utilized as nurseries and 1999; Larsson et al., 2000). feeding grounds in some species (Vögler et al., 2008). They are ambush The low diversity of macroinvertebrates in the carbonate-rich beds predators, preying predominantly on bony fish and at Åsen compared to nearby deposits at Ignaberga and Ivö Klack might (Baremore et al., 2010). Siversson et al. (2016) noted that the teeth of relate to both taphonomic processes (e.g., differential aragonite dis- the large-sized S. fortemordeo (most teeth of this species were extracted solution) and the composition of the original ecosystem (e.g., offshore from the Acutostrea-rich part of Bed 3) have a heavily worn cusp (see neritic settings with a scarcity of suitable hard substrates in the sandy- Siversson et al., 2016, figs. 6–8). They speculated that this may be the bottom environments versus rocky shoreline or carbonate shoal set- result of frequently striking the calcified rostra of belemnite . In tings: Surlyk and Sørensen, 2010; Sørensen and Surlyk, 2011). The contrast, teeth of the smaller but more abundant S. lundegreni are less abundance of pelagic belemnite rostra and scarcity of other in- worn (see Siversson et al., 2016, figs. 3–4). The latter species is re- vertebrate remains suggests that the oyster community that developed presented by teeth of highly variable sizes, indicating the presence of in attachment to rafted wood and small pebbles in Bed 3 represented a juvenile and adult individuals. The much larger-toothed S. fortemordeo specialized local ecosystem in which hard substrates for encrusting is represented mainly by large teeth, almost certainly from mature in- organisms were at a premium. dividuals (considering they are extremely large for the genus; see Siversson et al., 2016, fig. 8f–j). There are no modern examples of angel 9. Conclusions sharks occurring in large numbers in mangroves. The disappearance of hybodont shark teeth (and dermal scutes of Numerous oysters (Acutostrea incurva) from the Belemnellocamax sturgeons; Fig. 9H) above Bed 2 indicates diminishing influence of mammillatus zone at Åsen in the Kristianstad Basin possess encrustation riverine discharge (Siversson et al., 2016). An increasing distance away moulds on their left valves with features characteristic of the distal from the river mouth, which was most likely located a few kilometres to portions of woody subaerial axes. Some are referable to forms assigned the north, does not in itself preclude deposition in a coastal, fringing by previous authors to Pinites nathorstii, a putative cupressacean conifer. mangrove environment. However, as outlined above, several aspects of The relatively narrow range and normal distribution in diameters of the elasmobranch fauna are difficult to explain in the context of an axes encrusted by shells suggests winnowing of woody debris of a se- intertidal mangrove setting. Large apex-predatory lamniform sharks lected size range into the oyster community by water currents or a appear to have been common in the area, particularly otodontid sharks. preference for substrate items of specific size by the initially mobile In the comparatively well studied Western Interior Seaway of North oyster larvae. America, otodontid material of Late Cretaceous age is primarily asso- The foraminiferal suite from sediment within oyster body cavities in ciated with deeper, offshore palaeoenvironments (e.g., the Smoky Hill Bed 3 at Åsen differs from modern mangrove-associated assemblages by Chalk of the ; see Shimada, 2007 and Siversson lacking siliceous agglutinated Foraminifera, having around 5% abun- et al., 2015). In contrast, teeth of Archaeolamna kopingensis are abun- dance and a diverse array of Lagenida, and incorporating relatively dant in nearshore sandstone facies (Case, 1978, 1987). The relative common triserial Buliminida, and a significant percentage of planktonic increase in teeth of otodontids coupled with a decrease in relative taxa. The diversity and representation of foraminiferal groups

111 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114 collectively favour an offshore inner neritic environment with a rela- flygmagnetiska kartan Kristianstad SO. Sveriges Geologiska Undersökning Af121. tively clean sandy substrate. 55–99. pp. 1–120. – Berkeley, A., Perry, C.T., Smithers, S.G., Horton, B.P., Cundy, A.B., 2009a. Foraminiferal The chondrichthyan faunas of Beds 1 3 indicate a transgressive biofacies across mangrove–mudflat environments at Cocoa Creek, North Queensland, sequence with maximum water depths reached during the formation of Australia. Mar. Geol. 263, 64–86. the Acutostrea-rich interval in Bed 3. The abundance of shed teeth from Berkeley, A., Perry, C.T., Smithers, S.G., 2009b. Taphonomic signatures and patterns of test degradation on topical, intertidal benthic Foraminifera. Mar. Micropaleontol. 73, large, apex-predatory sharks and the numerical dominance of teeth of 148–163. bullhead and angel sharks are not compatible with the interpretation of Bonde, S.D., Kumaran, K.P.N., 2002. A permineralized species of mangrove fern a mangrove environment in the area during deposition of Bed 3. High Acrostichum L. from Deccan Intertrappean beds of India. Rev. Palaeobot. Palynol. – primary productivity and at least seasonally turbid waters are indicated 120, 285 299. Brotzen, F., 1936. Foraminiferen aus dem Schwedischen untersten Senon von Eriksdal in by the location of the site at the southern end of a drowned river valley Schonen. Årsb. Sver. Geol. Unders., Ser. C. 396. pp. 1–206. and the abundance of teeth from large, pelagic sharks. Dense popula- Brotzen, F., 1942. Die Foraminiferengattung Gavelinella nov. gen. und die systematik der – tions of benthic, thin-shelled invertebrates, for which there are no, or Rotaliformes. Årsb. Sver. Geol. Unders. ser. C. 451. pp. 1 50. Cappetta, H., Case, G.R., 1975. Contribution à l'étude des sélaciens du groupe Monmouth poor, fossil records in Bed 3, are indicated by the numerous collected (Campanien-Maestrichtien) du New Jersey. Palaeontogr. Abt. A 151, 1–46. teeth from durophagous elasmobranchs, in particular bullhead sharks Carlsson, J.G., 1938. AW Malms samling av kritfossil från Kristianstadsomradet. 1. but also the large shovelnose ray Rhinobatos casieri. Cephalopoda, , Lamellibranchiata, och Brachiopoda. Göteborgs Kungliga Vetenskaps och Vitterhets-Samhälles Handlingar. Femte följden. Ser. B. Bd. 6. pp. The coarse-grained component of Bed 3 (bivalve shells and be- 1–25 (5). lemnite rostra) relates to essentially autochthonous growth of an oyster Case, G.R., 1978. A new selachian fauna from the (Campanian) of community and a regular fall-out of guards from pelagic belemnites. Montana. Palaeontogr. Abt. A 160, 176–205. fi Case, G.R., 1987. A new selachian fauna from the late Campanian of Wyoming (Teapot The matrix of the bed itself is relatively ne grained (e.g., compared Sandstone Member, Mesaverde Formation, Big Horn Basin). Palaeontogr. Abt. A 197, with Beds 1 and 2), and the Foraminifera show minimal abrasion, in- 1–37. dicating relatively quiet-water conditions. The collective palaeontolo- Casier, E., 1947. Constitution et evolution de la racine dentaire des Euselachii, II. Étude comparative des types. Bull. Mus. Roy. d'Hist. Natur. Belgique. 23. pp. 1–32. gical and taphonomic evidence points to a neritic, but not nearshore or Castro, J.I., 1993. The shark nursery of Bulls Bay, South Carolina, with a review of the mangrove-like, depositional setting for the oyster-rich Bed 3 at Åsen. shark nurseries of the southeastern coast of the United States. Environ. Biol. Fish 38, Rather than being cemented to mangrove roots or stems, those oysters 37–48. – in Bed 3 associated with plant remains appear to have been attached to Christensen, W.K., 1969. Ivö Klack ecological and stratigraphical observations. Meddelelser fra Dansk Geologisk Forening 19, 135–138. small detrital subaerial axis fragments, some of which were clearly Christensen, W.K., 1975. Upper Cretaceous belemnites from the Kristianstad area in coniferous twigs with leaf scars. Other oysters were attached to pebbles . Fossils Strata 7, 1–69. and invertebrate shells signifying opportunistic use of any isolated hard Christensen, W.K., 1997. The Late Cretaceous belemnite family Belemnitellidae: tax- onomy and evolutionary history. Bull. Geol. Soc. Den. 44, 59–88. objects as substrates. The absence of evidence for mangrove commu- Collinson, M.E., Boulter, M.C., Holmes, P.R., 1993. Magnoliophyta (“Angiospermae”). In: nities at Åsen is consistent with the development of modern mangrove Benton, M.J. (Ed.), The Fossil Record. 2. Chapman & Hall, London, pp. 809–841. taxa much later in the Maastrichtian and Cenozoic based on the global Compagno, L.J.V., 1984. FAO species catalogue. In: Sharks of the World. An Annotated and Illustrated Catalogue of Sharks Species Known to Date 4. FAO Fisheries Synopsis, palynological record. pp. 250–655 ((125), 4, part 1, 1–249, part 2). Compagno, L.J.V., 2001. Sharks of the world. In: An annotated and illustrated catalogue Acknowledgments of shark species known to date. Bullhead, Mackerel and Carpet Sharks (Heterodontiformes, and Orectolobiformes), vol. 2 Food and Agriculture Organisation of the United Nations, Rome (FAO Species Catalogue for Financial support to Stephen McLoughlin by the Swedish Research Fishery Purposes. No. 1, Vol. 2). (269 pp.). Council (VR grant 2014-5234); and National Science Foundation Conwentz, H., 1892. Untersuchungen über fossile Hölzer Schwedens. Kongl. Sv. Vetensk.- (project #1636625) is gratefully acknowledged. Excavations at Åsen in Akad. Handl. 24. pp. 1–99. – Cook, T.D., Newbrey, M.G., Murray, A.M., Wilson, M.V.H., Shimada, K., Takeuchi, G.T., 2010 2012 were supported by grants to EE from the Royal Stewart, J.D., 2011. A partial skeleton of the Late Cretaceous lamniform shark, Physiographical Society of Lund. We gratefully acknowledge Prof. Vivi Archaeolamna kopingensis, from the Pierre Shale of western Kansas, U.S.A. J. Vertebr. Vajda and funding from LUCCI (Lund University Centre for studies of Paleontol. 31, 8–21. Cope, E.D., 1869. On the reptilian orders, Pythonomorpha and Streptosauria. Proc. Carbon Cycle and Climate Interactions) for supporting additional Boston Soc. Nat. Hist. 12, 250–266. fieldwork in Skåne. Linus Vajda is thanked for collection and collation Cushman, J.A., Valentine, W.W., 1930. Shallow water foraminifera from the channel is- of data on oyster shells. We thank three reviewers for their constructive lands of southern California. In: Contributions from the Department of Geology of Stanford University. 1. pp. 1–51. comments on the manuscript. Davis, J.W., 1890. On the fossil fish of the Cretaceous formations of . Sci. Trans. Roy. Dublin Soc. 4, 363–434. References Lamarck, J.B.P.A. de M. de, 1819. Histoire Naturelle des Animaux sans Vertèbres 6 (1). 1. Déterville, Paris, pp. 1–232. de Montfort, D., 1808. Conchyliologie systèmatique et classification méthodique des co- Abu Hamad, A.M.B., Amireh, B., El Atfy, H., Jasper, A., Uhl, D., 2016. Fire in a Weichselia- quilles; offrant leurs figures, leur arrangement générique, leurs descriptions dominated coastal ecosystem from the Lower Cretaceous (Barremian) of the Kurnub caractéristiques, leur noms; ainsi que leur synonymie en plusieurs langues. Group in NW Jordan. Cretac. Res. 66, 82–93. Conchyliologie systèmatique, Tome premiere. F. Schoell, Paris. (409 pp). Agassiz, J.L.R., 1833–1844. Recherches sur les poissons fossiles. 3 Petitpierre (Neuchâtel Dicken, M.L., Smale, M.J., Booth, A.J., 2006. Spatial and seasonal distribution of the et Soleure). (i–xxxii + 390 pp.). ragged-toothed shark Carcharias taurus along the coast of South Africa. Afr. J. Mar. Alvin, K.L., 1971. Weichselia reticulata (Stokes et Webb) Fontaine from the Wealden of Sci. 28, 603–616. Belgium. Mém. Inst. Roy. Sci. Nat. Belgique 166, 1–33. d'Orbigny, A., 1840. Memoire sur les Foraminiferes de la craie blanche du Bassin de Paris. Baremore, I.E., Murie, D.J., Carlson, J.K., 2010. Seasonal and size-related differences in Soc. Géol. Fr. Mém. 4, 1–51. diet of the Atlantic angel shark Squatina dumeril in the northeastern Gulf of Mexico. Ehrenberg, C.G., 1839. Uber die Bildung der Kreidefelsen und des Kreidemergels durch Aquat. Biol. 8, 125–136. unsichtbare Organismen. Abh. K. Akad. Wiss. Berlin. 1838. pp. 59. Barry, M.D., 1981. Distribution and ecology of oysters, Ostrea edulis (L.) in Kilkieran and Eichler, P.P.B., Rodrigues, A.R., Pereira, E.R.M., Eichler, B.B., Kahn, A., Vital, H., 2015. Bertraghboy Bays, Connemara, Co. Galway. Ir. Fish. Investig. B (Mar.) 24, 1–18. Foraminifera as environmental condition indicators in Todos os Santos Bay (Bahia, Bazzi, M., Einarsson, E., Newbrey, M.G., Campione, N.E., Kear, B.P., Siversson, M., 2017. Brazil). Open J. Ecol. 2015 (5), 326–342. A large Squalicorax (: Lamniformes) from the Late Cretaceous of Einarsson, E., Praszkier, A., Vajda, V., 2016. First evidence of the Cretaceous decapod Sweden and the association of sharks and mosasaurs. In: Society of Vertebrate Protocallianassa from Sweden. In: Kear, B.P., Lindgren, J., Hurum, J.H., 77th Annual Meeting, Program and Abstracts, pp. 79. Milàn, J., Vajda, V. (Eds.), Mesozoic Biotas of Scandinavia and its Arctic Territories. Beissel, I., 1891. Die Foraminiferen der Aachener Kreide. Abhandlungen der Preußischen 434. Geological Society, London, Special Publications, pp. 241–250. Geologischen Landesanstalt, Berlin. new ser. 3. pp. 1–76. Eklund, H., Friis, E.M., Pedersen, K.R., 1997. Chloranthaceous floral structures from the Belford, D.J., 1960. Upper Cretaceous foraminifera from the Toolonga Calcilutite and Late Cretaceous of Sweden. Plant Syst. Evol. 207, 13–42. Gingin Chalk, Western Australia. Bull. Bur. Mineral Resour. Geol. Geophys. 57, Ellison, A.M., Farnsworth, E.J., Merkt, R.E., 1999. Origins of mangrove ecosystems and 1–198. the mangrove biodiversity anomaly. Glob. Ecol. Biogeogr. 8, 95–115. Bergström, J., Sundquist, B., 1978. Kritberggrunden. In: Kornfält, K.-A., Bergström, J., El-Saadawi, W., Osman, R., El-Faramawi, M.W., Bkhat, H., Kamal-El-Din, M., 2016. On Carserud, L., Henkel, H., Sundquist, B. (Eds.), Beskrivning till berggrundskartan och the Cretaceous mangroves of Bahariya Oasis, Egypt. Taeckholmia. 36. pp. 1–16.

112 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

El-Soughier, M.I., Mehrotra, R.C., Zhoua, Z.-Y., Shi, G.-L., 2011. Nypa fruits and seeds data from great white sharks and makos. Geol. Soc. Am. Abstr. Programs 48, 163 from the Maastrichtian–Danian sediments of Bir Abu Minqar, South Western Desert, (paper No.). Egypt. Palaeoworld 20, 75–83. Legare, B., Kneebone, J., DeAngelis, B., Skomal, G., 2015. The spatiotemporal dynamics Erlström, M., Gabrielson, J., 1986. The Upper Cretaceous clastic deposits of Ullstorp, of habitat use by blacktip (Carcharhinus limbatus) and lemon (Negaprion brevirostris) Kristianstad basin, Scania. Geol. Fören. Stockh. Förh. 107, 241–254. sharks in nurseries of St. John, United States Virgin Islands. Mar. Biol. 162, 699–716. Erlström, M., Gabrielson, J., 1992. Petrology, fossil composition and depositional history Lidmar-Bergström, K., 1982. Pre-Quaternary geomorphological evolution in southern of the Ignaberga , Kristianstad Basin, Scania. Sver. Geol. Unders. 80, 1–30. Fennoscandia. Sver. Geol. Unders. C 785, 1–202. Fajemila, O.T., Langer, M.R., Lipps, J.H., 2015. Spatial patterns in the distribution, di- Link, H.F., 1807. Beschreibung der Naturalien-Sammlung der Universität zu Rostock. versity and abundance of benthic Foraminifera around Moorea (Society Archipelago, Vierte Abteilung: Fossile Ueberbleibsel organischer Körper, sogenannte French Polynesia). PLoS One 10 (12), e0145752. Versteinerungen. Rostock. (30 pp.). Falcon-Lang, H.J., 2004. A new anatomically preserved ginkgoalean genus from the Loeblich, A.R., Tappan, H., 1987. Foraminiferal Genera and Their Classification. Van Upper Cretaceous (Cenomanian) of Peclnov Quarry, near Prague, Czech Republic. Nostrand Reinhold, New York (970 pp.). (+ 847 pl). Palaeontology 47, 349–366. Loeblich, A.R., Tappan, H., 1992. Present status of foraminiferal classification. In: Falcon-Lang, H.J., 2005. Small cordaitalean trees in a marine-influenced coastal habitat Takayanagi, Y., Saito, T. (Eds.), Studies in Benthic Foraminifera. Tokai University in the Pennsylvanian Joggins Formation, Nova Scotia. J. Geol. Soc. 162, 485–500. Press, Tokyo, pp. 93–102. Farnsworth, E.J., Ellison, A.M., 1996. Sun-shade adaptability of the red mangrove, Lundegren, A., 1934. Kristianstadsområdet Kritbildningar. Geol. Fören. Stockh. Förh. 56, Rhizophora mangle (Rhizophoraceae): changes through ontogeny at several levels of 125–313. biological organization. Amer. J. Bot. 83, 1131–1143. Lundgren, B., 1876. Om belemniterna i sandkalken i Skåne. Ofversigt af Kongl. Vetensk.- Friis, E.M., 1984. Preliminary report of Upper Cretaceous angiosperm reproductive or- Akad. Forhandl. 1876. pp. 15–21. gans from Sweden and their level of organization. Ann. Mo. Bot. Gard. 71, 403–418. Lundgren, B., 1894. Jämförelse mellan molluskfaunan i mammillatus och mucronata Friis, E.M., 1990. Silvianthemum suecicum gen. et sp. nov., a new saxifragalean flower from zonerna i nordöstra Skåne (Kristianstadsområdet). Kongl. Sv. Vetensk.-Akad. Handl. the Late Cretaceous of Sweden. Biologiske Skrifter. Det Kongelige Danske 26. pp. 3–58 (6). Videnskabernes Selskab 36, 1–35. Malapris, M., 1965. Les et formes affinées du gisement Albien de Friis, E.M., Skarby, A., 1981. Structural preserved angiosperm flowers from the Upper Courcelles (Aube). Rev. Micropaleontol. 8, 131–150. Cretaceous of Sweden. Nature 291, 485–486. Manire, C.A., Gruber, S.H., 1993. A Preliminary Report on the Rate of Natural Mortality Friis, E.M., Crane, P.R., Pedersen, K.R., 2011. The Early Flowers and Angiosperm of Juvenile Lemon Sharks, Negaprion brevirostris. 115. NOAA Tech. Rep. NMFS, pp. Evolution. Cambridge University Press, Cambridge (600 pp.). 65–71. Gabrielson, J., 1991. Paleoecology of Late Cretaceous Shallow-water Benthic Márquez-Aliaga, A., Jiménez-Jiménez, A.P., Checa, A.G., Hagdorn, H., 2005. Early oysters Foraminifera from the Kristianstad Area, Scania, Southern Sweden. Licentiate Thesis. and their supposed Permian ancestors. Palaeogeogr. Palaeoclimatol. Palaeoecol. 229, Department of Geology, Lund University. 127–136. Gale, A.S., Sørensen, A.M., 2015a. Origin of the balanomorph barnacles (Crustacea, Marsson, T., 1878. Die Foraminiferen der weissen Schreibkreide der Insel Ruegen. Cirripedia, Thoracica): new evidence from the Late Cretaceous (Campanian) of Mitteilungen des naturwissenschaftlichen Vereins Neu-Vorpommern und Ruegen in Sweden. J. Syst. Palaeontol. 13, 791–824. Greifswald. 10. pp. 115–196. Gale, A.S., Sørensen, A.M., 2015b. and palaeoecology of thoracican cirripedes Mazer, S.J., Tiffney, B.H., 1982. Fruits of Wetherellia and Palaeowetherellia (Crustacea) from a Campanian rocky shoreline at Ivö Klack, southern Sweden. Cretac. (?Euphorbiaceae) from sediments in Virginia and Maryland. Brittonia 34, Res. 54, 212–242. 300–333. Gardunho, D.C.L., Gomes, C.P., Tagliaro, C.H., Beasley, C.R., 2012. Settlement of an McLaughlin, R.H., O'Gower, A.K., 1971. Life history and underwater studies of a het- unidentified oyster (Crassostrea) and other epibenthos on plastic substrates at a erodont shark. Ecol. Monogr. 41, 271–289. northern Brazilian mangrove island. Braz. J. Aquat. Sci. Technol. 16, 41–51. McLoughlin, S., 1993. Plant fossil distributions in some Australian Permian non-marine Gee, C.T., 2001. The mangrove palm Nypa in the geologic past of the New World. Wetl. sediments. Sediment. Geol. 85, 601–619. Ecol. Manag. 9, 181–194. McLoughlin, S., 2011. Glossopteris – insights into the architecture and relationships of an Geinitz, H.B., 1840. Charakteristik der Schichten und Petrefacten des sächsischen iconic Permian Gondwanan plant. J. Bot. Soc. Bengal 65, 93–106. Kreidegebirges, Zweites Heft. A. Das Land zwischen dem Plauen'schen Grunde bei Moberg, J.C., 1884. Cephalopoderaa i Sveriges kritsystem. I. Sveriges kritsystem sys- Dresden und Dohna. B. Fische, Crustaceen, Mollusken. – I – II. Arnold, Dresden, tematiskt framstaldt. Sveriges geologiska Undersokning, Serie C. 63. pp. 1–45. 31–62, taf. 9–16. Murray, J.W., 1976. A method of determining proximity of marginal seas to an ocean. Grönwall, K.A., 1915. Nordöstra Skånes kaolin− och kritbildningar samt deras praktiska Mar. Geol. 22, 103–119. användning. Sver. Geolog. Undersök. C. 261. pp. 1–185. Murray, J.W., 1991. Ecology and Palaeoecology of Benthic Foraminifera. Longman Gruber, S.H., DeMarignac, J.R.C., Hoenig, J.M., 2001. Survival of juvenile lemon sharks Scientific and Technical, New York (397 pp.). at Bimini, Bahamas, estimated by mark-depletion experiments. Trans. Am. Fish. Soc. Murray, J.W., 2006. Ecology and Applications of Benthic Foraminifera. Cambridge 130, 376–384. University Press, Cambridge (426 pp.). Halamski, A.T., Kvaček, J., Vajda, V., 2016. Late Cretaceous (Campanian) leaf and pa- Nelson, T.C., 1924. The attachment of oyster larvae. Biol. Bull. 46, 143–151. lynoflora from southern Skåne, Sweden. In: Kear, B.P., Lindgren, J., Hurum, J.H., Nilsson, S., 1827. Petrifica suecana Formationis Cretaceae, Descripta et Iconibus Milàn, J., Vajda, V. (Eds.), Mesozoic Biotas of Scandinavia and its Arctic Territories. Illustrata. Pars prior, Vertebrata et Sistens. Londini Gothorum, ex Officina 434. Geological Society, London, Special Publications, pp. 207–230. Berlingiana. (39 pp). Heupel, M.R., Hueter, R.E., 2002. Importance of prey density in relation to the movement Norling, E., 1981. Upper Jurassic and Lower Cretaceous geology of Sweden. Geol. Fören. pattern of juvenile blacktip sharks (Carcharhinus limbatus) within a coastal nursery Stockh. Förh. 103, 253–269. area. Mar. Freshw. Res. 53, 543–550. Norling, E., Skoglund, R., 1977. Der Sudwestrand Osteuropeischen Tafel im Bereich Iqbal, F.J., 2013. Paleoecology and of the Upper Cretaceous (Campanian), Schwedens. Z. Angew. Geol. 23, 449–458. Marine Strata at Åsen, Kristianstad Basin, Southern Sweden. MSc Thesis. 373. Nykvist, N., 1957. Kretaceiska vedrester vid Åsen i Skåne. Svenska Skogvårdsföreningens Department of Geology, Lund University, pp. 1–54. Tidskrift. 55. pp. 477–481. Jafar, S.A., Kar, R.K., 1996. Glossopterid flora: a case for wetland mangrove vegetation? Olsson, R.K., Nyong, E., 1984. A paleoslope model for Campanian–lower Maastrichtian In: Ninth International Gondwana Symposium, Hyderabad, January 1994. AAP foraminifera of New Jersey and Delaware. J. Foram. Res. 14, 50–68. Balkema, Rotterdam, pp. 331–337. Pessagno, E.A., 1967. Upper Cretaceous planktonic foraminifera from the west coastal Jeyabal, G., Muralidharan, S., Ramasamy, S., Rajeshwara Rao, N., Krishna Kumar, S., plain. Palaeontogr. Am. 5, 259–441. Parthasarathy, P., 2016. Vertical distribution of foraminifera species and palaeo- Philippe, M., 1992. Une mangrove à conifères au Jurassique en Franche-Comté. Bulletin ecology of the Pichavaram mangrove ecosystem, southeast coast of India. J. Coast. de l'Euforbe. 5. pp. 10–12. Sci. 3, 1–7. Plaziat, J.C., Cavagnetto, C., Koeniguer, J.C., Baltzer, F., 2001. History and biogeography Kathiresan, K., Bingham, B.L., 2001. Biology of mangroves and mangrove ecosystems. of the mangrove ecosystem, based on a critical reassessment of the paleontological Adv. Mar. Biol. 40, 81–251. record. Wetl. Ecol. Manag. 9, 161–180. Kent, D.V., Irving, E., 2010. Influence of inclination error in sedimentary rocks on the Pufahl, P.K., James, N.P., 2006. Monospecific oyster buildups, Murray Basin, Triassic and Jurassic apparent pole wander path for North America and implications South Australia: brackish water end member of the reef spectrum. Palaeogeogr. for cordilleran tectonics. J. Geophys. Res. Solid Earth 115 B10103. https://doi.org/ Palaeoclimatol. Palaeoecol. 233, 11–33. 10.1029/2009JB007205. Rafinesque, C.S., 1810. Caratteri di alcuni nuovi generi e nuove specie di animali e piante Krassilov, V.A., Zakharov, Yu.D., 1975. Pleuromeia from the Lower Triassic of the far east della Sicilia con varie osservazioni sopra i medesimi. San Filippo, Palermo. (105 pp.). of the U.S.S.R. Rev. Palaeobot. Palynol. 19, 221–232. Raymond, A., Phillips, T.L., 1983. Evidence for an Upper mangrove com- Kumpas, M.G., 1980. Seismic stratigraphy and tectonics in Hanö Bay, southern Baltic. munity. In: Teas, H.J. (Ed.), Tasks for Vegetation Science. Biology and Ecology of Stockh. Contrib. Geol. 34, 37–168. Mangroves 8. Junk, Boston, MA, pp. 19–30. Kvaček, J., Falcon-Lang, H.J., Daškova, J., 2005. A new Late Cretaceous ginkgoalean Rees, J., 1999. Late Cretaceous hybodont sharks from the Kristianstad Basin, southern reproductive structure Nehvizdyella gen. nov. from the Czech Republic and its whole- Sweden. N. Jahrb. Geol. Paläontol., Monatshefte 1999, 257–270. plant reconstruction. Amer. J. Bot. 92, 1958–1969. Retallack, G., 1975. The life and times of a Triassic lycopod. Alcheringa 1, 3–29. Larsson, K., Solakius, N., Vajda-Santivanez, V., 2000. Foraminifera and palynomorphs Retallack, G.J., 1977. Reconstructing Triassic vegetation of eastern Australasia: a new from the greensand-limestone sequences (Aptian–Coniacian) in southwestern approach for the biostratigraphy of Gondwanaland. Alcheringa 1, 247–277. Sweden. N. Jb. Geol. Paläont. (Abh.) 216, 277–312. Reuss, A.E., 1845. Die Versteinerungen der böhmischen Kreideformation. Abtheilung 1, Last, P.R., Stevens, J.D., 2009. Sharks and Rays of Australia. CSIRO Publishing (656 pp.). Teil 2. E. Schweizerbart, Stuttgart. pp. 1–148. Leder, R.M., Perez, V.J., Badaut, T., 2016. Simplifying the methods – body length esti- Reuss, A.E., 1862. Palaeontologische Beitràge. Sitzungsber. Kais. Akad. Wiss. Wien Math.- mates for Carcharocles using associated tooth sets and jaw width related Naturwiss. KL. 44. pp. 301–342.

113 S. McLoughlin et al. Palaeogeography, Palaeoclimatology, Palaeoecology 498 (2018) 99–114

Revets, S.A., 1996. The Generic Revision of Five Families of Rotaliine Foraminifera. 34. and surrounding areas. Norges Geol. Unders. Bull. 316, 289–311. Cushman Foundation for Foraminiferal Research, Special Publication, pp. 1–113. Springer, S., 1967. Social organization of shark populations. In: Gilbert, P.W., Satyanarayana, B., Husain, M.-L., Ibrahim, R., Ibrahim, S., Dahdouh-Guebas, F., 2014. Mathewson, R.F., Rall, D.P. (Eds.), Sharks, Skates and Rays. Johns Hopkins University Foraminiferal distribution and association patterns in the mangrove sediments of Press, Baltimore, pp. 149–174. Kapar and Matang west peninsular Malaysia. J. Sustain. Sci. Manag. 9, 32–48. Srinivasan, V., Friis, E.M., 1989. Taxodiaceous conifers from the Upper Cretaceous of Schönenberger, J., Friis, E.M., 2001. Fossil flowers of ericalean s.l. affinity from the Late Sweden. In: Biologiske Skrifter, Det Kongelige Danske Videnskabernes Selskab. 35. Cretaceous of southern Sweden. Amer. J. Bot. 88, 467–480. pp. 1–57. Schrank, E., 1987. Palaeozoic and Mesozoic palynomorphs from northeast Africa (Egypt Stephenson, L.W., 1916. Coelenterata. In: Clark, W.B. (Ed.), Systematic Paleontology of and Sudan) with special reference to Late Cretaceous pollen and dinoflagellates. Berl. the Upper Cretaceous Deposits of Maryland. 752–757. Maryland Geological Survey, geowissensch. Abh., R. A. 75. pp. 249–310. Baltimore, Maryland, pp. 48–49. Schweitzer, C.E., Lacovara, K.J., Smith, J.B., Lamanna, M.C., Lyon, M.A., Attia, Y., 2003. Surlyk, F., Sørensen, A.M., 2010. An early Campanian rocky shore at Ivö Klack, southern Mangrove-dwelling crabs (: Brachyura: Necrocarcinidae) associated with Sweden. Cretac. Res. 31, 567–576. from the Upper Cretaceous (Cenomanian) of Egypt. J. Paleontol. 77, Thanikaimoni, G., 1987. Mangrove palynology. Institut français de Pondichéry, Travaux 888–894. de la Section Scientifique et Technique. 24. pp. 1–100. Segura-Zarzosa, J.C., Abitia-Cardenas, L.A., Galvan-Magana, F., 1997. Observations on Thibault, N., Husson, D., Harlou, R., Gardin, S., Galbrun, B., Huret, E., Minoletti, F., 2012. the feeding habits of the shark Heterodontus francisci Girard 1854 (Chondrichthyes: Astronomical calibration of upper Campanian–Maastrichtian carbon isotope events Heterodontidae), in Laguna San Ignacio, Baja California Sur, Mexico. Ciencias and calcareous plankton biostratigraphy in the Indian Ocean (ODP hole 762C): im- Marinas. 23. pp. 111–128. plication for the age of the Campanian–Maastrichtian boundary. Palaeogeogr. Sen, A., Ghosh, M., Khanderao, P., Das, S.K., Roy Chowdhury, D., Sarkar, P.K., Saha, R., Palaeoclimatol. Palaeoecol. 337, 52–71. Bhadury, P., 2016. Modern benthic foraminiferal assemblages from the world's lar- Todd, R., Brönnimann, P., 1957. Recent Foraminifera and Thecamoebina from the eastern gest deltaic mangrove ecosystem, the Sundarbans. Mar. Biodivers. 46, 421–431. Gulf of Paria, Trinidad. J. Foram. Res. 3, 1–43. Sender, L.M., Villanueva-Amadoz, U., Pons, D., Diez, J.B., García-Ávila, M., Ferrer, J., Tomlinson, P.B., 1986. The Botany of Mangroves. Cambridge University Press, Cambridge 2015. New reconstruction of Weichselia reticulata (Stokes et Webb) Fontaine in Ward (413 pp.). emend. Alvin, 1971 based on fertile remains from the middle of Spain. Hist. Vajda, V., 2001. Aalenian to Cenomanian terrestrial palynofloras of SW Scania, Sweden. Biol. 27, 460–468. Acta Palaeontol. Pol. 46, 403–426. Shimada, K., 2007. Skeletal and dental anatomy of lamniform shark, Cretalamna appen- Vajda, V., Wigforss-Lange, J., 2006. The Jurassic-Cretaceous transition of southern diculata, from Upper Cretaceous Niobrara Chalk of Kansas. J. Vertebr. Paleontol. 27, Sweden—palynological and sedimentological interpretation. Prog. Nat. Sci. 16, 1–38. 584–602. Vajda-Santivanez, V., Solakius, N., 1999. Palynomorphs, foraminifera, and calcispheres Silantieva, N., Krassilov, V., 2006. Weichselia Stiehler from Lower Cretaceous of Makhtesh from the greensand–limestone transition at Arnager, Bornholm: evidence of trans- Ramon, Israel: new morphological interpretation and taxonomical affinities. Acta gression during the late Cenomanian to early Coniacian. GFF 121, 281–286. Palaeobotanica. 46 (119eobo). Vakhrameev, V.A., 1991. Jurassic and Cretaceous Floras and Climates of the Earth. Siverson, M., 1992. Biology, dental morphology and taxonomy of lamniform sharks from Cambridge University Press, Cambridge, UK, pp. 318. the Campanian of the Kristianstad Basin, Sweden. Palaeontology 35, 519–554. van der Ham, R.W.J.M., van Konijnenburg-van Cittert, J.H.A., Kieft, B.N., Walsmit Sachs, Siverson, M., 1993. Late Cretaceous and Danian Neoselachians from Southern Sweden. A., 2011. Mosacaulis spinifer gen. et sp. nov.: an enigmatic Maastrichtian plant. Rev. 110. Lund Publications in Geology, pp. 1–28. Palaeobot. Palynol. 168, 51–67. Siversson, M., Lindgren, J., Newbrey, M.G., Cederström, P., Cook, T.D., 2015. van der Ham, R.W.J.M., Konijnenburg-van Cittert, J.H.A., Jagt, J.W.M., Indeherberge, L., Cenomanian–Campanian (Late Cretaceous) mid-palaeolatitude sharks of Cretalamna Meuris, R., Deckers, M.J.M., Renkens, S., Laffineur, J., 2017. Seagrass stems with appendiculata type. Acta Palaeontol. Pol. 60, 339–384. attached roots from the type area of the Maastrichtian stage (NE Belgium, SE Siversson, M., Cook, T.D., Cederström, P., Ryan, H.E., 2016. Early Campanian (Late Netherlands): morphology, anatomy, and ecological aspects. Rev. Palaeobot. Palynol. Cretaceous) squatiniform and synechodontiform selachians from the Åsen locality, 241, 49–69. Kristianstad Basin, Sweden. In: Kear, B.P., Lindgren, J., Hurum, J.H., Milàn, J., Vajda, van Hinsbergen, D.J., de Groot, L.V., van Schaik, S.J., Spakman, W., Bijl, P.K., Sluijs, A., V. (Eds.), Mesozoic Biotas of Scandinavia and its Arctic Territories. 434. Geological Langereis, C.G., Brinkhuis, H., 2015. A paleolatitude calculator for paleoclimate Society, London, Special Publications, pp. 251–275. studies. PLoS One 10 (6), e0126946. Skarby, A., 1968. Extratriporopollenites (Pflug) emend. from the Upper Cretaceous of Vögler, R., Milessi, A.C., Quiñones, R.A., 2008. Influence of environmental variables on Scania, Sweden. In: Stockholm Contributions in Geology. 16 Almqvist & Wiksell, the distribution of Squatina guggenheim (Chondrichthyes, Squatinidae) in the Stockholm (60 pp.). Argentine-Uruguayan Common Fishing Zone. Fish. Res. 91, 212–221. Skarby, A., 1986. Normapolles anthers from the Upper Cretaceous of southern Sweden. Voigt, S., Gale, A.S., Jung, C., Jenkyns, H.C., 2012. Global correlation of upper Rev. Palaeobot. Palynol. 46, 235–256. Campanian–Maastrichtian successions using carbon-isotope stratigraphy: develop- Skarby, A., Morbelli, M.A., Rowley, J.R., 2004. Fine structure of Rossipollis reticulatus from ment of a new Maastrichtian timescale. Newsl. Stratigr. 45, 25–53. the Upper Cretaceous of southern Sweden. Can. J. Bot. 82, 927–937. Wallace, R.K., 2001. Cultivating the Eastern Oyster, Crassostrea virginica. Southern Sopher, D., Erlström, M., Bell, N., Juhlin, C., 2016. The structure and stratigraphy of the Regional Aquaculture Centre Publication 432, 1–4. sedimentary succession in the Swedish sector of the Baltic Basin: new insights from White, M.P., 1928. Some index Foraminifera of the Tampico Embayment area of Mexico. vintage 2D marine seismic data. Tectonophysics 676, 90–111. Part I. J. Paleontol. 2, 280–317. Sørensen, A.M., Surlyk, F., 2008. A brachiopod boring (Podichnus) in a Late Cretaceous White, W.T., Potter, I.C., 2004. Habitat partitioning among four elasmobranch species in oyster from a mangrove-like environment, Skåne, Sweden. Lethaia 41, 295–298. nearshore, shallow waters of a subtropical embayment in Western Australia. Mar. Sørensen, A.M., Surlyk, F., 2011. Taphonomy and palaeoecology of the gastropod fauna Biol. 145, 1023–1032. from a Late Cretaceous rocky shore, Sweden. Cretac. Res. 32, 472–479. Wiman, C., 1920. Some reptiles from the Niobrara Group in Kansas. Geol. Inst. Uppsala Spjeldnæs, N., 1975. Palaeogeography and facies distribution in the Tertiary of Denmark Bull. 18, 9–18.

114