You are on page 1of 82

 

 
Sedimentology of Hirnantian glaciomarine deposits in the Balkan Terrane,
western Bulgaria: Fixing a piece of the north peri-Gondwana jigsaw puzzle

Athanas Chatalov

PII: S0037-0738(17)30016-7
DOI: doi:10.1016/j.sedgeo.2017.01.004
Reference: SEDGEO 5154

To appear in: Sedimentary Geology

Received date: 23 November 2016


Revised date: 13 January 2017
Accepted date: 18 January 2017

Please cite this article as: Chatalov, Athanas, Sedimentology of Hirnantian glaciomarine
deposits in the Balkan Terrane, western Bulgaria: Fixing a piece of the north peri-
Gondwana jigsaw puzzle, Sedimentary Geology (2017), doi:10.1016/j.sedgeo.2017.01.004

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT
1

1 Sedimentology of Hirnantian glaciomarine deposits in the Balkan Terrane, western

2 Bulgaria: fixing a piece of the north peri-Gondwana jigsaw puzzle

T
4 Athanas Chatalov

IP
5

R
6 Department of Mineralogy, Petrology and Economic Geology, Faculty of Geology and Geography,

SC
7 Sofia University St. Kliment Ohridski, 15 Tsar Osvoboditel Bd., BG-1504 Sofia, Bulgaria

8 Tel: +359887732049

NU
9 Fax: +35929446487

10 E-mail address: chatalov@gea.uni-sofia.bg


MA
11
D
P TE
CE
AC
ACCEPTED MANUSCRIPT
2

12 ABSTRACT

13 Glaciomarine deposits of late Hirnantian age in the western part of the Palaeozoic Balkan Terrane

14 have persistent thickness (~7 m) and lateral uniformity in rock colour, bedding pattern, lithology, and

15 sedimentary structures. Four lithofacies are distinguished from base to top: lonestone-bearing

T
IP
16 diamictites, interbedded structureless mudstones, crudely laminated diamictites, and finely laminated

17 mudstones. The diamictites are clast-poor to clast-rich comprising muddy to sandy varieties. Their

R
18 compositional maturity is evidenced by the very high amount of detrital quartz compared to the

SC
19 paucity of feldspar and unstable lithic grains. Other textural components include extraclasts derived

NU
20 from the local Ordovician basement, mudstone intraclasts, and sediment aggregates. Turbate

21 structures, grain lineations, and soft sediment deformation of the matrix below larger grains are locally
MA
22 observed. Sedimentological analysis reveals that deposition occurred in an ice-intermediate to ice-

23 distal, poorly agitated shelf environment by material supplied from meltwater buoyant plumes and

24 rain-out from ice-rafted debris. Remobilization by mass-flow processes (cohesive debris flows and
D

25
TE

slumps) was an important mechanism particularly for the formation of massive diamictites. The

26 glaciomarine deposits represent a typical deglaciation sequence reflecting retreat of the ice front
P

27 (grounded or floating ice sheet), relative sea-level rise and gradually reduced sedimentation rate with
CE

28 increasing contribution from suspension fallout. This sequence was deposited on the non-glaciated

29 shelf of the intracratonic North Gondwana platform along the southern margin of the Rheic Ocean.
AC

30 The Hirnantian strata of the Balkan Terrane can be correlated with similar glaciomarine deposits

31 known from peri-Gondwana terranes elsewhere in Europe showing clear ‘Armorican affinity’. Several

32 lines of evidence suggest that the provenance of siliciclastic material was associated mainly with

33 sedimentary recycling of mature sands which had been deposited across North Gondwana in Cambrian

34 and pre-glacial Ordovician times.

35

36 Keywords: diamictite; ice-distal; deglaciation sequence; provenance; palaeogeography.

37

38

39
ACCEPTED MANUSCRIPT
3

40 1. Introduction

41

42 The recognition of ancient glaciogenic deposits is of fundamental importance to

T
43 palaeoclimatic, palaeogeographic, palaeoenvironmental and palaeotectonic reconstructions

IP
44 (Chumakov, 1985; Crowell, 1999; Menzies, 2002; Hambrey and Glasser, 2003; Eyles and

R
45 Januszczak, 2004; Eyles, 2008). It provides information about the major controls, duration

SC
46 and severity of cold climatic conditions, size and dynamics of the ice masses, related eustatic

47 changes, the complexity of glaciated basins, depositional settings and products, the relation to

48
NU
adjacent non-glacial environments, and the impact of ancient glaciations on the evolution of
MA
49 life on Earth. The sedimentary records from pre-Pleistocene glaciations are dominated by

50 glaciomarine deposits which are typically thicker than their terrestrial counterparts (Eyles and

51 Eyles, 1992; Eyles, 1993) being well documented from all major glacial episodes in the
D
TE

52 Precambrian and Phanerozoic (Hambrey and Harland, 1981; Deynoux et al., 1994; Eyles,

53 2008). Glaciomarine sediments have a higher potential for preservation because of sufficient
P

54 accommodation space and also because they are covered rapidly by transgressive deposits due
CE

55 to the postglacial sea-level rise (Eyles, 1993). Glacially influenced marine environments
AC

56 provide valuable records of ice sheet fluctuations because they occur beyond the limits of ice

57 advance and erosion (Licht, 2009). Glaciomarine deposits may also be potential reservoir

58 facies and targets for hydrocarbon exploration in sedimentary basins (Potter et al., 1995;

59 Fielding et al., 2012; Lang et al., 2012, and references therein).

60 The Hirnantian age of the late Ordovician was a time of extensive ice-sheet growth over

61 large parts of the Gondwana Supercontinent which occupied a southern polar position

62 throughout most of the early Palaeozoic (Ghienne et al., 2007; Le Heron and Dowdeswell,

63 2009; Le Heron and Howard, 2010; Torsvik and Cocks, 2011; Pohl et al., 2016). Various

64 glacial non-depositional features and glacial rocks have been recognized in many countries,
ACCEPTED MANUSCRIPT
4

65 e.g., from Africa, Europe, Asia, and South America (Beuf et al., 1971; Young et al., 2004;

66 Turner et al., 2005; Ghienne et al., 2007; Schönian and Egenhoff, 2007; Gutiérrez-Marco et

67 al., 2010; Le Heron et al., 2010; Ghienne, 2011, among others). Glaciomarine deposits have

T
68 been identified in particular from glaciated and non-glaciated shelf areas of the intracratonic

IP
69 North Gondwana platform (Ghienne et al., 2007). They consist of diverse ice-contact, ice-

R
70 proximal and ice-distal facies which provide information about advance/retreat phases of the

SC
71 ice sheet(s) as well as the palaeogeographic position of many Gondwana and peri-Gondwana

72 terranes during the late Ordovician (e.g., Steiner and Falk, 1981; Robardet and Doré, 1988;

73
NU
Brenchley and Štorch, 1989; Leone et al., 2002; Monod et al., 2003; Denis et al., 2007; Le
MA
74 Heron et al., 2007; Ghavidel-syooki et al., 2011; Moreau, 2011; Schönlaub et al., 2011; Couto

75 et al., 2013). Prolific glaciogenic hydrocarbon systems occur in the Upper Ordovician/Lower

76 Silurian of North Africa and the Middle East (Le Heron et al., 2009; Huuse et al., 2012, and
D
TE

77 references therein) with postglacial black shales being important source rocks (Lüning et al.,

78 2000).
P

79 This study focuses on the sedimentology of Hirnantian glaciomarine strata which are
CE

80 exposed in the western part of the Palaeozoic Balkan Terrane in western Bulgaria. A detailed
AC

81 petrographic description of the siliciclastic rocks is presented with special emphasis on the

82 depositional characteristics of the diamictites. The field and microscopic observations are

83 used to interpret the particular glaciomarine environment and to clarify the relative

84 contribution of various processes to the formation of different lithofacies. The changing

85 proximity of the ice front and related temporal changes in sea level, sedimentation rate,

86 depositional mechanisms and products are also discussed. Conclusions are drawn about the

87 provenance of glacially supplied siliciclastic material on the basis of microtextural features

88 and major-element chemistry. A correlation with similar glaciomarine deposits known from
ACCEPTED MANUSCRIPT
5

89 other peri-Gondwana terranes and results from previous studies are integrated to determine

90 the palaeogeographic position and affinity of the Balkan Terrane during the late Ordovician.

91

T
92 2. The Ordovician System in Bulgaria – a brief synopsis

IP
93

R
94 The occurrence of Ordovician rocks in Bulgaria was identified by Haberfelner and

SC
95 Bončev (1934) who found trilobite and graptolite fauna in Palaeozoic exposures from the

96 western part of the country. Later, palaeontological and biostratigraphical studies resulted in

97
NU
the subdivision of Ordovician series and stages as well as establishing formal and informal
MA
98 lithostratigraphic units (Spassov, 1960, 1968; Sachanski and Tenchov, 1993; Yanev et al.,

99 1995; Gutierrez-Marco et al., 2003; Sachanski and Tanatsiev, 2010, among others).

100 Three Palaeozoic terranes (Moesian, Balkan and Thracian) were distinguished in the
D
TE

101 basement of the eastern Balkan Peninsula (Fig. 1) with Gondwana and peri-Gondwana origins

102 although each terrane is characterized by specific tectonic and sedimentary evolution (Yanev,
P

103 1993, 1997, 2000; Haydoutov and Yanev, 1997; Yanev et al., 2005). Palaeomagnetic data for
CE

104 the Balkan Terrane obtained from outcrops in neighbouring Serbia (Milicevic, 1994)
AC

105 indicated palaeolatitudes of 29–50°S, 30–40°S and 38°S during the Early, Middle and Late

106 Ordovician, respectively. The occurrence of Middle to Upper Ordovician oolitic ironstones

107 and glaciomarine diamictites (Sachanski, 1994; Yanev, 2000; Gutierrez-Marco et al., 2003)

108 was considered as good evidence for deposition in the temperate humid zone at about 40°S

109 (Yanev et al., 2005). In the Ordovician period the Balkan Terrane represented a shallow

110 marine environment characterized by siliciclastic deposition (Yanev et al., 2005, 2006). It was

111 also presumed that the Balkan Terrane was closely linked to the ‘Armorican Terrane

112 Assemblage’, i.e., during its entire Palaeozoic evolution (Yanev et al., 2005, 2006;

113 Winchester et al., 2006). The origin and palaeogeographic affinity of the Moesian Terrane
ACCEPTED MANUSCRIPT
6

114 during the early Palaeozoic remains unclear as different interpretations were proposed for its

115 western part, i.e., in Bulgaria and Romania (e.g., Yanev et al., 2006; Oczlon et al., 2007;

116 Kalvoda and Bábek, 2010). Geochronological and geochemical studies on the metamorphic

T
117 basement of the Thracian Terrane, i.e., present-day Rhodope and Serbo-Macedonian massifs,

IP
118 revealed the presence of North Gondwana and peri-Gondwana protoliths of early Palaeozoic

R
119 age (Himmerkus et al., 2009; Bonev et al., 2013; Antić et al., 2016). According to a recently

SC
120 published global reconstruction for the late Ordovician, Moesia was an integral part of the

121 Armorica sub-terrane and the Rhodope massif belonged to the intra-Alpine sub-terrane of the

122
NU
ribbon-like Galatian superterrane (Stampfli et al., 2013). Palaeoclimatic indicators and
MA
123 palaeomagnetic data support a continuous migration of the Balkan and Moesian Terranes

124 from temperate southern latitude in the Ordovician/Silurian to the northern arid zone in the

125 Permain (Yanev, 1993; Haydoutov and Yanev, 1997; Yanev, 2000; Yanev et al., 2005).
D
TE

126 According to Stampfli et al. (2013) the Galatian superterrane became detached in segments

127 from Gondwana during the Devonian, starting with the Armorica sub-terrane around 400 Ma.
P

128 The thickest Ordovician sequence in Bulgaria is known from the Balkan Terrane,
CE

129 reaching up to 4700 m in the western part of the country (Yanev, 2000). It comprises about
AC

130 2000 m thick marine sedimentary strata (Yanev et al., 2006) including predominantly

131 siliciclastic rocks (claystones, siltstones, sandstones), minor chert, and rare oolitic ironstones.

132 The Middle to Late Ordovician age of these deposits was determined on the basis of

133 macrofossils (graptolites, trilobites) and palynomorphs (Spassov, 1960; Kalvacheva, 1990;

134 Sachanski, 1993; Gutiérez-Marco et al., 2002, 2003; Lakova and Sachanski, 2004). Various

135 taxa and associations of acritarchs, trilobites and brachiopods from the western Balkan

136 Terrane are closely related to peri-Gondwana provinces (Kalvacheva, 1990; Gutierrez-Marco

137 et al., 2002, 2003).

138
ACCEPTED MANUSCRIPT
7

139 3. Tectonic and stratigraphic framework

140

141 Ordovician sedimentary rocks are mainly exposed in the western Balkan Terrane.

T
142 Geographically, they occur in the western sector of Stara planina Mountains which are

IP
143 integral part of the Alpine Carpathian-Balkan belt (Fig. 2). In terms of Alpine tectonic

R
144 subdivision the Palaeozoic outcrops belong to the Svoge Unit which is the northernmost

SC
145 allochthone fragment of the Srednogorie Zone (Ivanov, 1988). The latter represents a specific

146 element of the so called Intermediate Balkanides (Fig. 2, inset map) and was formed during

147
NU
the late Cretaceous on a variegated Hercynian basement and a Lower Mesozoic orogenic
MA
148 edifice. After the Middle Eocene the Srednogorie moved northwards and was thrusted over

149 parts of the West Balkan Zone (i.e., External Balkanides) as a result of general compression.

150 From the south the Svoge Unit is bounded by other tectonic units of the Srednogorie Zone and
D
TE

151 the superimposed Neogene-Quaternary Sofia graben. The pre-Mesozoic basement of the

152 Svoge Unit was greatly influenced by Hercynian and Alpine deformations.
P

153 The Ordovician succession outcropping in the western part of Srednogorie Zone
CE

154 consists of five lithostratigraphic units (Angelov et al., 2010, 2011): siltstone-argillite
AC

155 metaformation, Grohoten Formation, Tseretsel Formation, Sirman Formation, and the

156 lowermost part of Saltar Formation. According to Sachanski and Tanatsiev (2010) the four

157 formal units have chronostratigraphic range from the Darriwilian Stage of the Middle

158 Ordovician Series to the Hirnantian Stage of the Upper Ordovician Series (Fig. 3). The

159 siliciclastic rocks studied herein are referred to as Sirman Fm (Sachanski and Tenchov, 1993).

160 Non-fosiliferous strata of the underlying Tseretsel Fm (Spassov, 1960) correspond to the

161 Katian Stage as defined by their stratigraphic position (Sachanski, 1994; Sachanski and

162 Tanatsiev, 2010). These light grey-greenish, fissile mudstones display a change of the rock

163 colour to dark grey in the uppermost 2-3 m of the unit, showing a sharp lithological contact
ACCEPTED MANUSCRIPT
8

164 (planar or slightly undulating) with diamictites of the Sirman Fm (Fig. 4a, b). Thin-bedded,

165 dark grey to black cherts (lydites), organic-rich shales and siliceous shales of the Saltar Fm

166 (Sachanski and Tenchov, 1993) have a sharp lithological but conformable contact with the

T
167 Sirman Fm (Fig. 4c). The chronostratigraphic range of the chert-shale succession was

IP
168 determined to span from the upper Hirnantian Stage of the Upper Ordovician Series to the

R
169 lower Telychian Stage of the Llandovery Series on the basis of graptolites (Sachanski, 1993;

SC
170 Lakova and Sachanski, 2004). In particular, the presence of Normalograptus persculptus

171 (now referred to as Metabolograptus persculptus, e.g., Melchin et al., 2011; Štorch and

172
NU
Schönlaub, 2012) at the very base of the Saltar Fm gives grounds to define a late Hirnantian
MA
173 age (persculptus Biozone) for the Sirman Fm (Sachanski, 1993).

174

175 4. Materials and methods


D
TE

176

177 Ten stratigraphic sections with exposed Upper Ordovician rocks were selected across
P

178 the study area for description and sampling in the field (Fig. 2). The thicknesses of Sirman Fm
CE

179 were measured and macroscopic characteristics of the rocks (stratal boundaries, colours,
AC

180 sedimentary structures, lithologies, clast morphology, composition and size) were described.

181 Seven polished slabs were prepared from laminated rock types and sixty-six samples (plus

182 three additional samples from the underlying Tseretsel Fm) were collected for observation

183 with transmitted light microscopy. All thin sections were stained with dilute HCl solution of

184 Alizarine red S and potassium ferricyanide to detect the presence of carbonates.

185 The bulk rock composition of diamictites and mudstones was determined by wet

186 chemical analysis of twelve powdered samples. To estimate the relative extent of

187 palaeoweathering and to examine the provenance composition the chemical index of

188 alteration (CIA, Nesbitt and Young, 1982) was calculated from data for the major oxides
ACCEPTED MANUSCRIPT
9

189 using molar proportions in the formula Al2O3/[Al2O3 + CaO* + Na2O + K2O] × 100. Because

190 CaO* represents CaO content in the silicate minerals only, corrections were made for CaO in

191 apatite using the available P2O5 data (Fedo et al., 1995). If the remaining number of moles

T
192 was less than that of Na2O, the CaO value was adopted as CaO*, and if the number of moles

IP
193 for CaO was greater than Na2O, the CaO* was assumed to be equivalent to Na2O (McLennan,

R
194 1993).

SC
195

196 5. Description of the Hirnantian siliciclastic rocks

197
NU
MA
198 5.1. Stratification, macropetrography and lithofacies

199

200 The lower part of the Sirman Fm consists of thick-bedded (0.6-0.8 m), dark grey or
D
TE

201 grey-greenish diamictites showing massive or uncommon slump structures (lithofacies 1) and

202 two layers (15 cm and 5 cm thick, respectively) of dark grey, structureless mudstones
P

203 (lithofacies 2) having sharp conformable contacts (Fig. 4d, e). The stacked diamictite beds
CE

204 also have sharp non-erosive basal contacts with local development of low-amplitude load
AC

205 casts. Subangular to well rounded extrabasinal clasts derived from light grey-greenish

206 mudstones (Fig. 5a, b) and light grey sandstones (Fig. 5c) display a maximum size of about 3

207 cm (i.e., outsized clasts), diverse shapes, and subvertical, random or subparallel orientation to

208 bedding. It is noteworthy that most mudstone extraclasts greatly resemble rocks of the

209 underlying Tseretsel Fm. A second clast type includes dark grey mudstones showing the same

210 maximum size, mostly flattened shape, various degree of roundness, and random orientation

211 (Fig. 5d). The conspicuous similarity of those clasts to the interbedded mudstones of

212 lithofacies 2 implies their intrabasinal origin. Upwards in the section the diamictites become

213 characterized by dominantly medium bedding (0.3-0.5 m) and crude parallel lamination
ACCEPTED MANUSCRIPT
10

214 (lithofacies 3) with the discrete laminae displaying various thicknesses (up to 5 mm), colours,

215 and grain sizes (Figs. 4f, 5e). These rocks contain the same types of extraclasts and intraclasts

216 but having maximum granule size, dominantly elongated shape, and subparallel orientation to

T
217 bedding. The siliciclastic succession is topped by a 0.5 m thick bed of grey, finely laminated

IP
218 mudstones (lithofacies 4) showing gradual transition from the underlying laminated

R
219 diamictites (Fig. 5f). The measured thickness of the Sirman Fm ranges between 6.30 m and

SC
220 7.20 m (Fig. 3).

221

222
NU
5.2. Rock classification and depositional textures
MA
223

224 Under the microscope, the two diamictite lithofacies can be classified more precisely by

225 applying the non-genetic textural scheme of poorly sorted sediments proposed by Moncrieff
D
TE

226 (1989) and later modified by Hambrey (1994). This classification is based on the proportion

227 between sand and mud (as matrix) against the relative amount of gravel clasts. The Hirnantian
P

228 rocks are defined as clast-poor to clast-rich diamictites because the gravel-sized component
CE

229 varies quantitatively in the 1-15% interval. Muddy (Fig. 6a, b), intermediate (Fig. 6c, d) and
AC

230 sandy (Fig. 6e, f) varieties of these diamictites are distinguished according to the sand/mud

231 ratio with poor sorting of the sand-sized material being a typical characteristic. The main

232 textural constituents are monocrystalline (strongly dominant) and polycrystalline (sparse)

233 detrital quartz grains as their total amount constantly exceeds 90% of the sand fraction. The

234 grain morphology is subangular to well rounded with moderate degree of sphericity being

235 most common. The bulk of monocrystalline grains are non-undulatory while the rest exhibit

236 sweeping extinction (Fig. 7a). Few quartz grains are characterized by deformation lamellae

237 (Fig. 7b). Some grains are recognized as having definite origin, for example, volcanic quartz

238 shows partly preserved idiomorphic crystal habit, rounded edges, embayments and uniform
ACCEPTED MANUSCRIPT
11

239 extinction (Fig. 7c), while reworked sedimentary quartz is distinguished by the presence of

240 abraded quartz overgrowths (Fig. 7d). Individual crystals in the polycrystalline grains have

241 various numbers, shapes and sizes, and their crystal boundaries are either sutured or straight

T
242 indicating metamorphic or plutonic igneous source (Fig. 7e, f). Exotic rock fragments in the

IP
243 diamictites include quartzites (Fig. 7g), quartzose and arkosic sandstones (Fig. 7h, k),

R
244 microcrystalline chert (Fig. 7i), and granitoids (Fig. 7j). Much more common are lithic grains

SC
245 derived from mudstones, as most of them resemble by colour and texture the rocks of the

246 underlying Tseretsel Fm. These mudstone extraclasts have diverse shapes and degree of

247
NU
roundness being invariably characterized by thin oxidized rims (Fig. 8a, b) while few grains
MA
248 contain quartz veinlets inherited from the parent rock (Fig. 8c). Other mudstone clasts imply

249 intrabasinal source by displaying good rounding, absence of sharply outlined boundaries or

250 oxidized rims, locally manifested effects of mechanical compaction, and great similarity to
D
TE

251 the interbedded mudstones of lithofacies 2 (Fig. 8d–f). A specific type of ‘intraclasts’

252 comprises sediment aggregates that bear a resemblance to the surrounding diamictite
P

253 groundmass. These aggregates rarely show elongated morphology and distinct boundaries
CE

254 (Fig. 8g) but more commonly have irregular shape with diffuse boundaries and higher amount
AC

255 of mud (Fig. 8h–j). Minor detrital constituents in the diamictites include feldspar grains (Figs.

256 6e, 7l) and muscovite flakes (Fig. 6b). The scarce heavy minerals are largely dominated by

257 zircon (Fig. 7f) with tourmaline and rutile being present as single grains. Besides the parallel

258 lamination which is defined by vertical changes in the colour, grain size, and/or sand/mud

259 ratio (Figs. 6f, 9i), some specific microfabrics are also observed in the rocks. For example, in

260 the diamictites of lithofacies 1 there are very rare and isolated occurrences of galaxy/turbate

261 structure, i.e., circular alignment of grains and matrix without a core grain (Fig. 9a, b) or

262 around a large clastic grain (Fig. 9c), as well as grain lineations, i.e., linear arrangement of

263 three or more grains with their long axes being parallel to one another (Fig. 6f). In the crudely
ACCEPTED MANUSCRIPT
12

264 laminated diamictites (lithofacies 3), elongated clasts or mineral grains display subvertical

265 orientation of their long axis (Figs. 7f, 9d), while soft deformation of the matrix occurs below

266 some detrital grains of very coarse sand or granule sizes (Fig. 9d, e).

T
267 The interbedded mudstones of lithofacies 2 are characterized by a homogeneous texture

IP
268 where the silt fraction occurs in minor or negligible amount (Fig. 9f, g). Similarly to the mud

R
269 matrix of the diamictites these rocks consist of low birefringent clay particles and randomly

SC
270 dispersed silt grains largely dominated by quartz. Laminae formed of quartz silt without

271 grading are typical features of the mudstones comprising lithofacies 4 (Fig. 9h). The thin

272
NU
laminae (0.2-0.5 mm) are planar to slightly undulating and only locally discontinuous without
MA
273 showing evidence of significant soft deformation or erosion of underlying units.

274

275 5.3. Diagenetic alteration and late deformation


D
TE

276

277 The major diagenetic alterations in the diamictites include compaction and mineral
P

278 replacement. The former is recognized as mechanical grain rearrangement, ductile


CE

279 deformation of mudstone intraclasts (Fig. 8b, e, f) or sediment agrregates (Fig. 8g–j), and
AC

280 scarce grain fracturing plus pressure solution effects which are only present in the sandy

281 diamictites (Fig. 6f). Very small amounts (<1%) of calcite replacing feldspar or lithoclasts are

282 detected in two samples while replacement of feldspar grains by clay minerals is more

283 common (Fig. 7l). Coarse pyrite crystals and aggregates display clear replacement pattern in

284 some crudely laminated diamictites (Fig. 5e). A spectacular feature in the same lithofacies is

285 the occurrence of chamosite strain fringes developed between quartz grains and/or quartz-rich

286 rock fragments (Fig. 9j), and quartz strain fringes grown around subhedral to euhedral pyrite

287 crystals (Fig. 9k). These fringes have been interpreted as products of coaxial progressive
ACCEPTED MANUSCRIPT
13

288 deformation under low temperature conditions and broadly related to the Variscan orogeny of

289 the Balkan Terrane (Chatalov, 2014).

290

T
291 6. Major element chemistry and CIA

IP
292

R
293 The major element concentrations of the Hirnantian diamictites and mudstones are

SC
294 given in Table 1. Analytical data show that most of the diamictite samples (n = 8) are rich in

295 SiO2 (69.13–81.84%). The source of silica is mainly detrital quartz but also includes quartz-

296
NU
bearing rock fragments, mudstone lithoclasts, and silicate minerals in the matrix. The
MA
297 diamictites have a narrow range of Al2O3 (8.10–13.68%) which is basically related to clay

298 minerals forming the matrix and mudstone lithoclasts. The relatively low content of CaO

299 (0.59–1.72%), Na2O (0.12–1.47%) and K2O (1.33–2.69%) is due to the minor presence of
D
TE

300 feldspars, muscovite, some lithic grains and probably illite in the matrix. The mudstone

301 samples (n = 4) have lower values for SiO2 (34.44–62.59%) but higher values for Al2O3
P

302 (8.89–20.88%) which is a consequence of the large amount of clay minerals and
CE

303 predetermines a lower SiO2/Al2O3 ratio compared to the diamictites. The slightly higher
AC

304 contents of CaO (0.65–3.02%), Na2O (0.53–1.82%) and K2O (0.83–4.41%) in these rocks are

305 likely associated with illite and smectite minerals (see Chatalov, 2014).

306 The calculated average CIA value is 68.8 ± 7.2 (n = 8) for the diamictites and 72.3 ± 5.3

307 (n = 4) for the mudstones, respectively. This slight difference is seemingly related to the

308 presence of detrital feldspar plus feldspar-rich lithic grains (i.e., derived from arkosic

309 sandstones and granitoids) and the less amount of aluminous clays in the coarse-grained

310 lithofacies. Such an effect of grain sorting on the mineralogical and chemical composition of

311 sediments and rocks, and consequently on the CIA values, is a well known relationship as

312 demonstrated by Nesbitt and Young (1996) and Nesbitt et al. (1996). All recorded values are
ACCEPTED MANUSCRIPT
14

313 close to the average shale (70-75) and in particular to the average composition (~70) of

314 Phanerozoic shale (Nesbitt and Young, 1982; Taylor and McLennan, 1985; Condie et al.,

315 2001).

T
316

IP
317 7. Interpretation of the results

R
318

SC
319 7.1. Evidence for glaciomarine sedimentation

320

321
NU
Diamictites are poorly sorted rocks that can be products of various geological settings
MA
322 regardless of latitude or climate, including the formation of glacial deposits, debrites,

323 turbidites, olistostromes, lahars, regoliths, submarine slumps, tectonites, and impact breccias

324 (Eyles and Januszczak, 2004). In glacial and glacially influenced environments diamictite can
D
TE

325 form from lodgement and deformation of subglacial sediments, in situ meltout of ice and

326 release of poorly sorted basal, englacial or supraglacial debris, slumping and reworking of
P

327 glaciogenic debris in proglacial, glaciolacustrine or glaciomarine environments, ice keel


CE

328 ‘ploughing’ by grounding ice masses, and from rainout of fine-grained sediment and ice-
AC

329 rafted debris (Arnaud and Etienne, 2011).

330 Some characteristics of the studied diamictites are typical for both glacial terrestrial and

331 glacial marine deposits (see Hambrey and Glasser, 2003), e.g., poor sorting of clasts,

332 variations of grain and clast morphology and composition, constant mix of clasts over wide

333 area, and presence of well preserved unstable lithic fragments. In addition, the rocks show

334 explicit evidence for a glaciomarine origin, i.e., sedimentation in a marine setting by

335 combination of glacial and marine processes. Thus, they correspond to several criteria

336 proposed for the identification of glaciomarine deposits (Hambrey and Glasser, 2003) which

337 precludes a glacial terrestrial origin: distinct stratification, lower and upper boundaries with
ACCEPTED MANUSCRIPT
15

338 marine strata, presence of lonestones, random arrangement of clasts, and evidence for

339 resedimentation by subaqueous gravity flows. Moreover, the Hirnantian strata occur as

340 laterally persistent deposits having almost constant thickness (i.e., sheet-like geometry) and

T
341 showing uniformity in the stratification pattern, rock colours, lithological composition,

IP
342 depositional structures and textures across the whole area of exposure (see Eyles et al., 1985).

R
343 In comparison, relatively thin (<10-15 m) tabular sheets of massive or stratified diamictites

SC
344 interbedded with mudstones and sandstones are characteristic of modern and late Cenozoic

345 cool-temperate or subpolar glaciomarine shelf sequences (e.g., Dowdeswell et al., 1998;

346
NU
Elverhøi and Henrich, 2002; Powell and Cooper, 2002), and glaciomarine successions having
MA
347 similar thicknesses and lithologies are also known from the pre-Mesozoic geological record

348 (Brenchley and Štorch, 1989; Ghienne, 2003; Halverson et al., 2004; Marenssi et al., 2005;

349 Dobrzinski and Bahlburg, 2007; Schönlaub et al., 2011; Ghienne et al., 2013, among others).
D
TE

350 Deposition in a glaciomarine environment is further emphasized by the sharp but non-

351 erosional basal contacts of the diamictite beds, sedimentary structures (i.e., parallel
P

352 lamination, slumps, low-amplitude load casts), sand-dominated grain size, interlayered marine
CE

353 mudstones, and presence of intraclasts derived from those mudstones.


AC

354 Other textural characteristics of the Hirnantian rocks are not diagnostic but appear

355 consistent with the interpreted glaciomarine origin, for example, the occurrence of cm-sized

356 lonestones (i.e., outsized clasts) in the non-laminated diamictites. According to Eyles and

357 Januszczak (2004) lonestones are not a reliable indicator of glacial environments because they

358 are commonly associated with sediment gravity flows. It has been also demonstrated that

359 outsized clasts in fine-grained strata can be formed by other specific processes not particularly

360 related to glaciation (Bennett et al., 1996; Tachibana, 2013, and references therein).

361 Lonestones are usually interpreted as ice-rafted debris, i.e., dropstones, if they deform the

362 underlying sediment (Thomas and Connell, 1985). In this case such deformation was not
ACCEPTED MANUSCRIPT
16

363 observed macroscopically but this may be due to relatively small size of the lonestones,

364 firmness of the substrate at the time of their settling (see Álvaro and Van Vliet-Lanoë, 2009),

365 redeposition by sediment gravity flows, later compaction effects, and/or massive (i.e., non-

T
366 laminated) appearance of the diamictites. Meanwhile, possible evidence for ice-rafted origin

IP
367 includes the subvertical orientation of some elongated lonestones, local concentrations of up

R
368 to three outsized clasts, and the overall clast lithology that partly differs from the local

SC
369 basement (i.e., mudstones of the Tseretsel Fm).

370 In the crudely laminated diamictites, syndepositional disturbance of the underlying

371
NU
matrix and draping of sediment on the top of very coarse sand- to granule-sized detrital grains
MA
372 suggest that the latter originated as ice-rafted debris. This small-scale soft sediment

373 deformation was presumably controlled by the particular shape and relatively small size of the

374 dropped grains (Thomas and Connell, 1985). The anisometric extraclasts and mineral grains
D
TE

375 with preserved steep dip of their long axis likewise indicate drop release from floating ice and

376 subsequent piercing into stiff or soft substrate (see Álvaro and Van Vliet-Lanoë, 2009). In
P

377 addition, the presence of glacial erratics, including unstable lithic fragments such as arkosic
CE

378 sandstones, granitoids, and some mudstones (not derived from the local basement), implies
AC

379 far-travelled source material supplied by ice rafting.

380 Lithified sediment aggregates that resemble the surrounding groundmass in diamictites

381 are similar to aggregates described in the literature termed sediment pellets that have been

382 related to various glacial processes (e.g., Piotrowski et al., 2006; Tomkins et al., 2009; Li et

383 al., 2011, Cowan et al., 2012, and references therein). The so called ‘till pellets’ (Ovenshine,

384 1970) can originate as unconsolidated sediment filling interstices in glacial ice and carried

385 seawards by icebergs to be released through the water column as cohesive pellets upon

386 melting (Ovenshine, 1970; Gilbert, 1990; Goldschmidt et al., 1992). The recognition of such

387 formerly frozen masses of ice-rafted debris in the studied diamictites is based on their diffuse
ACCEPTED MANUSCRIPT
17

388 boundaries (Arnaud and Etienne, 2011), predominantly irregular shape, and lower sand/mud

389 ratio with respect to the diamictite groundmass. Thus, they can be distinguished from the

390 more rarely observed sediment aggregates (pellets) that display distinct boundaries, mostly

T
391 elongated morphology, very similar texture to the surrounding diamictite, and probably

IP
392 represent intrabasinal rip-up clasts.

R
393 The presence of galaxy/turbate structure is mainly known from debris flows (e.g.,

SC
394 Menzies and Zaniewski, 2003; Phillips, 2006; Isbell et al., 2016) and subglacial tills (e.g.,

395 Denis et al., 2010; Cowan et al., 2012; Lea and Palmer, 2014). However, this rotational

396
NU
structure has been also documented from Neogene and Quaternary glaciomarine diamicts
MA
397 (Carr et al., 2000; Nelson et al., 2009; Kilfeather et al. 2010). In this case, the rare occurrence

398 of galaxy/turbate structure in the diamictites not associated with other micro-scale

399 deformation features (see Menzies, 2000) supports the interpretation of a glaciomarine origin
D
TE

400 for the diamictites (e.g., Carr, 2001).

401 To sum up, the combination of macro- and micropetrographic characteristics of the
P

402 Upper Ordovician rocks shows close correspondence to other ancient glaciomarine deposits
CE

403 (Fortuin, 1984; Eyles et al., 1985; Štorch, 1990; Kellerhals and Matter, 2003; Allen et al.,
AC

404 2004; Bernárdez et al., 2006; Dobrzinski and Bahlburg, 2007; Álvaro and Van Vliet-Lanoë,

405 2009; Ampaiwan et al., 2009; Perez Loinaze et al., 2010; Ghienne et al., 2010; Couto et al.,

406 2013, among many others). Furthermore, indirect evidence comes from the regional

407 correlation with Hirnantian glaciogenic successions exposed in some countries of Europe,

408 Africa, and Asia.

409

410 7.2. Depositional environment and processes

411
ACCEPTED MANUSCRIPT
18

412 Glaciomarine sedimentation is controlled by basin geometry, water depth, relief of the

413 sea floor, hydrodynamic regime, volume of meltwater input, debris amount in glacier ice and

414 icebergs, and the proximity of a glacial front (Eyles et al., 1985). Glaciomarine environments

T
415 may be subdivided into ice-marginal to ice-proximal, i.e., dominated by glacial processes, and

IP
416 ice-intermediate to ice-distal, i.e., glacially influenced or dominated by non-glacial marine

R
417 processes (Eyles and Eyles, 1992). The combination of several characteristics of the

SC
418 Hirnantian strata suggests ice-intermediate to ice-distal origin, i.e., deposition from several to

419 tens of kilometers away from the ice margin. These diagnostic features include: relatively

420
NU
small and persistent thickness of the glaciomarine succession (i.e., sheet-like geometry);
MA
421 uniformity with respect to the stratification pattern (i.e., distinct bedding), overall lithology

422 (i.e., dominant diamictites and subordinate mudstones) and sedimentary structures; sandy

423 (i.e., gravel-poor) character of the diamictites; presence of cm-sized lonestones; homogeneous
D
TE

424 composition of the lithic fragments and intraclasts (see Eyles et al., 1985; Štorch, 1990;

425 Kellerhals and Matter, 2003; Allen et al., 2004; Dobrzinski and Bahlburg, 2007; Álvaro and
P

426 Van Vliet-Lanoë, 2009; Ampaiwan et al., 2009; Ghienne et al., 2010; Perez Loinaze et al.,
CE

427 2010; Arnaud and Etienne, 2011; Couto et al., 2013). In this context, equally important is the
AC

428 virtual absence of ice-marginal or ice-proximal facies. The Hirnantian strata lack diagnostic

429 features such as coarse-grained (i.e., gravel-rich) deposits, rapid lateral or vertical changes in

430 thickness and lithofacies, irregular bed geometries, major erosional surfaces, fining-up and

431 coarsening-up units, large-scale foreset bedding, intensive gravity-induced resedimentation,

432 and glaciotectonic deformation (e.g., Lønne, 1995; Allen et al., 2004; Rieu et al., 2006;

433 Ampaiwan et al., 2009; Henry et al., 2010; Isbell, 2010; Passchier and Erukanure, 2010; Perez

434 Loinaze et al., 2010; Arnaud and Etienne, 2011; Arnaud, 2012; Busfield and Le Heron, 2013;

435 Couto et al., 2013; Koch and Isbell, 2013; Le Heron et al., 2014; Girard et al., 2015).

436 Laminated sediments may form in both ice-proximal and ice-distal modern (e.g., Ó Cofaigh
ACCEPTED MANUSCRIPT
19

437 and Dowdeswell, 2001) and ancient environments (e.g., Couto et al., 2013). However, some

438 characteristics of the laminated lithofacies 3 and 4 are more consistent with ice-distal origin,

439 i.e., sand- or mud-dominated grain size, horizontal orientation and undeformed pattern of

T
440 laminae, and lack of outsized clasts. Furthermore, the complete absence of high-energy

IP
441 depositional or erosional structures in conjunction with the non-winnowed textures and low

R
442 degree of sorting of the detrital material in all diamictites and mudstones suggests

SC
443 sedimentation in poorly agitated mid- to outer-shelf setting, i.e., most probably below the

444 storm-wave base.

445
NU
Sediment material is supplied in glaciomarine systems by meltwater buoyant plumes
MA
446 released from the glacial margin, ice-rafted debris, mass-flow events, and ocean currents

447 (Eyles et al., 1985; Eyles, 1993). The Hirnantian primary diamicts originated mainly from

448 coupled suspension settling of clay- to sand-sized particles from turbid meltwater plumes and
D
TE

449 rain-out of fine- to coarse-grained ice-rafted debris. This general conclusion is based on the

450 following evidence: a) blanket-like geometry of the glaciomarine deposits draping the
P

451 substrate; b) lack of grading and wave- or current-induced depositional and erosional
CE

452 structures; c) random grain alignment and textural variation (i.e., from mud- to sand-
AC

453 dominated) in the diamictites of lithofacies 1; and d) planar lamination with subparallel

454 orientation of elongated grains and clasts in the diamictites of lithofacies 3. The turbid plumes

455 most likely emanated from the front of a distant grounded or floating glacier while the ice-

456 rafted debris may have been derived from far-travelled icebergs or seasonal sea ice.

457 Distinction between the latter two sources is difficult in ancient successions especially for

458 sand-sized material (Gilbert, 1990; Goldschmidt et al., 1992; Álvaro and Van Vliet-Lanoë,

459 2009) and requires detailed study of grain microtextures (e.g., Stickley et al., 2009; St John et

460 al., 2015). Nevertheless, the diamictite lithology, variable roundness of the detrital material,

461 presence of diverse exotic clasts not derived from the local basement (for comparison see
ACCEPTED MANUSCRIPT
20

462 Isbell et al., 2013) and the occurrence of ‘till pellets’ imply far-travelled iceberg-related

463 transport in an ice-distal depositional environment. The crudely laminated structure of

464 diamictites forming the upper part of the Hirnantian succession probably reflects fluctuations

T
465 in meltwater discharge from the distant grounded/floating glacier and/or changes in the ratio

IP
466 between suspension fallout and ice-rafting.

R
467 A subaqueous mass flow origin for the primary diamicts can be ruled out considering

SC
468 the sheet-like geometry, presence of ice-rafted debris and parallel lamination, scarcity of soft-

469 sediment deformation, no correlation between the bed thickness and the maximum clast size,

470
NU
and lack of associated facies unambiguously produced by sediment gravity flows (e.g.,
MA
471 turbidites). However, redeposition of the thick-bedded diamictites by such flows, i.e.,

472 cohesive debris flows (Mulder and Alexander, 2001; Talling et al., 2012), is implied by their

473 stacked nature (less than 2 m thick beds), sharp non-erosive basal contacts (probably as a
D
TE

474 result from hydroplaning of individual flows), lack of internal structures (except slumped

475 beds), grading and preferred orientation of the clasts, formation of load casts, very poor
P

476 sorting, presence of extraclasts derived from basement lithologies, locally abundant mud
CE

477 matrix and mudstone intraclasts, and absence of disturbance below lonestones (see López-
AC

478 Gamundí, 1991; Visser, 1994; Mulder and Alexander, 2001; Eyles and Januszczak, 2004;

479 Arnaud and Eyles, 2006; Dobrzinski and Bahlburg, 2007; Ampaiwan et al., 2009; Henry et

480 al., 2010, 2012; Arnaud and Etienne, 2011; Couto et al., 2013; Isbell et al., 2016). In the same

481 context, it is noteworthy that some of the above described textural characteristics, e.g.,

482 outsized clasts, galaxy/turbate structures and sediment aggregates (pellets), are known from

483 deposits of subaqueous debris flows. Moreover, although grain lineations are most common

484 within subglacial tills (van der Meer, 1996; Hiemstra and Rijsdijk, 2003; Cowan et al., 2012),

485 they have been also interpreted as related to debris flow processes (Harris, 1998; Menzies and

486 Zaniewski, 2003).


ACCEPTED MANUSCRIPT
21

487 Subaqueous glaciogenic debris flows may have run-out distances of hundreds of

488 kilometres, even on shallow slopes (Dowdeswell et al., 1998; Laberg and Vorren, 2000;

489 Taylor et al., 2002). Therefore, while the Hirnantian diamictites are believed to have

T
490 originally formed by dominant rain-out and suspension settling, those having massive or

IP
491 slump structures were particularly remobilized by mass-flow processes (Eyles, 1993; Arnaud

R
492 and Eyles, 2006) as a result from high sedimentation rates, local oversteepened slopes with

SC
493 unstable water-saturated substrate, seismic shocks, storm events, meltwater surges, impact of

494 calving icebergs events, or combination of these processes (Boulton, 1990; Álvaro and Van

495
NU
Vliet-Lanoë, 2009; Arnaud and Etienne, 2011). Similarly, the coeval operation of fallout
MA
496 processes and sediment gravity flows in glaciomarine (including ice-distal) environments has

497 been interpreted or presumed in many studies of ancient diamictites (e.g., López-Gamundí,

498 1991; Visser, 1997; Marenssi et al., 2005; Arnaud and Eyles, 2006; Rieu et al., 2006;
D
TE

499 Dobrzinski and Bahlburg, 2007; Le Heron et al., 2007, 2014; Álvaro and Van Vliet-Lanoë,

500 2009; Ampaiwan et al., 2009; Perez Loinaze et al., 2010; Lang et al., 2012; Henry et al.,
P

501 2012; Couto et al., 2013; Deschamps et al., 2013; Dineen et al., 2013). In summary, the
CE

502 balance between transported material by meltwater turbid plumes, rain-out from ice-rafted
AC

503 debris and resedimentation by gravity flows could have varied but the homogeneous

504 sedimentation ultimately resulted in a regionally extensive, blanket-like glaciomarine facies.

505 Episodic bottom currents may have further assisted in the distribution of glaciogenic detrital

506 material across the shelf despite the lack of preserved current-formed structures in the

507 diamictites. Slight variations in the total thickness of the glaciomarine succession were

508 probably caused by uneven sea floor topography (e.g., Eyles et al., 1985), laterally variable

509 supply of siliciclastics from meltwater plumes, and/or locally enhanced redeposition by

510 sediment gravity flows.


ACCEPTED MANUSCRIPT
22

511 The sharp and conformable contacts between the diamictites and interbedded mudstones

512 record abrupt changes in the energy regime and sediment supply. The structureless mudstones

513 may have formed by suspension settling of dominantly clay-sized particles supplied from

T
514 sand-free meltwater plumes during quiescent periods of suppressed ice-rafting (i.e.,

IP
515 Dowdeswell et al., 2000; Ó Cofaigh and Dowdeswell, 2001). Alternatively, their deposition

R
516 occurred as a result of mud density flows (sensu Talling et al., 2012) producing so called

SC
517 densite mud (see also Isbell et al., 2016). While the former mechanism applies for the finely

518 laminated mudstones forming the top of the glaciomarine succession, it is likely that subtle

519
NU
changes in the turbid meltwater discharge and coeval background hemipelagic sedimentation
MA
520 (i.e., fine terrigenous deposition unrelated to meltwater plumes) produced laminated ungraded

521 sediments in an open-marine, low-energy setting during the maximum glacial retreat.

522
D
TE

523 7.3. Implications for a deglaciation sequence

524
P

525 Glaciomarine environments have complex dynamics which are controlled by ice
CE

526 advances or retreats and related sea-level changes. In general, most glaciomarine
AC

527 sedimentation takes place during glacial retreats when large volumes of meltwater are

528 released from grounded or floating ice sheets, and hence, sediments deposited during

529 deglaciation phases constitute the bulk of the stratigraphic record (Eyles and Eyles, 1992,

530 2010). Such deposits have great potential for preservation in marine environments because

531 they are covered rapidly by transgressive deposits due to postglacial sea-level rise, and also

532 because of the sufficient accommodation space (Eyles, 1993).

533 The Katian mudstones of the Tseretsel Fm forming the immediate pre-Hirnantian

534 basement have been interpreted as deep marine deposits exclusively on the basis of trace

535 fossil analysis (Aceñolaza and Yanev, 2001). Their sharp lithological contact with the
ACCEPTED MANUSCRIPT
23

536 overlying Sirman Fm may be the only record of distant ice sheet advance and related sea-level

537 fall, while the glaciomarine deposits were accumulated mainly during the following glacial

538 retreat and sea-level rise (Fig. 10). Only the thin mudstone interbeds within the lower

T
539 diamictite interval probably reflect minor re-advances or stabilisation of the ice front between

IP
540 phases of intensive glacier ice melting. This general pattern of glacially influenced

R
541 sedimentation resulted in the formation of a typical deglaciation sequence (sensu Visser,

SC
542 1997). Several lines of evidence suggest increasing distance of the depositional locus from the

543 ice front with progressive change in the style of glaciomarine sedimentation and resultant

544
NU
deposits. For example, the upsection transition from dominantly structureless to crudely
MA
545 laminated diamictites is accompanied by decreasing bed thickness, disappearance of

546 lonestones, slightly increasing mud content, and diminishing amount and size of mudstone

547 intraclasts and extraclasts derived from the local basement. Moreover, macro- and micro-
D
TE

548 characteristics indicative of mass-flow processes, e.g., slumps, load casts, random orientation

549 of clasts, galaxy/turbate structures, grain lineations and intrabasinal rip-up clasts from
P

550 diamictites (but not ‘till pellets’), are lacking in the upper part of the glaciomarine succession.
CE

551 The clear evidence for dropped ice-rafted grains and the appearance of parallel lamination
AC

552 implies dominant fallout deposition in a more ice-distal setting probably influenced by

553 episodic bottom currents but not dominated by sediment gravity flows. Further decrease of the

554 glacial influence is inferred from the gradual upward transition from crudely laminated, clast-

555 poor diamictites to finely laminated, clast-free mudstones. The greatly diminished thickness

556 and grain size of the discrete laminae reflect reduced sedimentation rate along with increasing

557 distance from the melting ice source by analogy with observations from modern glaciomarine

558 settings (e.g., Boulton, 1990; Ó Cofaigh and Dowdeswell, 2001, and references therein). The

559 formation of fine lamination is particularly interpreted to indicate the increasing contribution

560 from background hemipelagic sedimentation in an open-marine, low-energy environment


ACCEPTED MANUSCRIPT
24

561 during maximum recession of the ice front (see Marenssi et al., 2005; Moreau, 2011;

562 Passchier et al., 2011) in the absence of ice-rafting and most likely intermittent supply of fine-

563 grained material from turbid meltwater plumes. The outlined retrogradational trend continues

T
564 in the uppermost Hirnantian to Llandovery strata of the Saltar Fm with the occurrence of

IP
565 postglacial sediments, e.g., black cherts (lydites) and siliceous shales. Their sharp and

R
566 conformable contact with the underlying glaciomarine strata suggests abrupt environmental

SC
567 change after the final glacial retreat (see Henry et al., 2012), and therefore these deposits were

568 formed during a major transgressive phase in a deeper marine setting (Yanev et al., 2005)

569
NU
without any glacial influence. Such rapid eustatic sea-level rise as a result of deglaciation is
MA
570 commonly related to the deposition of black shales as evidenced from the Cryogenian and

571 Lower Palaeozoic sedimentary records (e.g., Lüning et al., 2000; Armstrong et al., 2005; Page

572 et al., 2007; Le Heron et al., 2009, 2011; Melchin et al., 2013).
D
TE

573 Similar deglaciation sequences reflecting glacial retreat with concomitant transition

574 from ice-proximal to ice-distal sedimentation were described from Phanerozoic glaciomarine
P

575 deposits by Visser (1997), Troedson and Riding (2002), Ghienne (2003), Marenssi et al.
CE

576 (2005), Vesely and Assine (2006), Ghienne et al. (2010), López-Gamundí (2010), Lang et al.
AC

577 (2012), and Dineen et al. (2013).

578

579 7.4. Palaeoweathering and provenance – clues from the CIA and micropetrography

580

581 The Chemical Index of Alteration (CIA) is the most widely used proxy to determine the

582 intensity of weathering and associated climatic regime in the source area (Bahlburg and

583 Dobrzinski, 2011). It measures quantitatively the extent to which primary minerals such as

584 feldspars are transformed into secondary aluminous clays with removal of mobile cations

585 (Ca2+, Na+, K+) which is enhanced during chemical weathering (Nesbitt and Young, 1982).
ACCEPTED MANUSCRIPT
25

586 While unweathered igneous rocks have CIA values below 55 reflecting cool and/or arid

587 conditions, intense weathering in humid and warm climate results in high concentrations of

588 residual kaolinite clays and/or Al oxihydroxides producing CIA values close to 100. The

T
589 average CIA values in shales (70-75) largely correspond to muscovite, illite and smectite

IP
590 compositions indicating a moderate degree of weathering. The CIA has been successfully

R
591 applied in a number of studies on Precambrian and Phanerozoic glacial deposits including

SC
592 glaciomarine mudstones and diamictites (Nesbitt and Young, 1982, 1996; Visser and Young,

593 1990; Young and Nesbitt, 1999; Young, 2001; Dobrzinski et al., 2004; Young et al., 2004;

594
NU
Scheffler et al., 2006; Rieu et al., 2007; Passchier and Erukanure, 2010; von Eynatten et al.,
MA
595 2012; Rose et al., 2013; Gaschnig et al., 2014; Rashid and Ganai, 2015; Ding et al., 2016; Li

596 et al., 2016). Because weathering products dominantly occur in the finer grain-size fractions,

597 CIA values of mudrocks are used to better assess the maximum weathering level in the source
D
TE

598 area (Nesbitt and Young, 1982).

599 The calculated CIA values for the Hirnantian diamictites (63–76) and mudstones (67–
P

600 76) indicate moderate degree of weathering in the source area. The lower mean values for the
CE

601 diamictites (68.8) relative to the mudstones (72.3) reflect less amount of aluminous clays in
AC

602 the rock matrix and the same grain size control on CIA has been recorded in other studies of

603 modern and ancient glacial deposits (e.g., Visser and Young, 1990; Nesbitt and Young, 1996;

604 Young and Nesbitt, 1999; von Eynatten et al., 2012). The obtained results conform to

605 previously performed XRD studies on mudstones of the Sirman Fm which revealed the

606 presence of partly ordered R=1 illite/smectite plus minor amounts of kaolinite (Chatalov,

607 2014).

608 A more informative method of representing the CIA than simple comparison of

609 numerical values employs a ternary diagram of A-CN-K with plotted molar proportions of

610 Al2O3, CaO*+Na2O and K2O (Nesbitt and Young, 1984). This triangle may indicate
ACCEPTED MANUSCRIPT
26

611 weathering trends, provenance composition, and post-depositional chemical modifications

612 (Fedo et al., 1995). In the A–CN–K diagram, data points of the diamictites and mudstones

613 outline a cluster that is oriented subparallel to the A–CN line (Fig. 11). Thus, it closely

T
614 follows the ideal weathering trend as would be predicted from both theoretical and empirical

IP
615 weathering studies (Nesbitt and Young, 1984). The recorded slight deviation from the ideal

R
616 weathering trend is probably due to minor post-depositional increase in K (Fedo et al., 1995)

SC
617 that could be linked to the presence of illite in the rock matrix and as alteration product of

618 feldspar because the other possible source, i.e., replacement of plagioclase by authigenic K-

619
NU
feldspar (Fedo et al., 1995), was not observed microscopically. In constraining the initial
MA
620 compositions of the parent rocks, the samples plot on a line that is consistent with a chemical

621 weathering trend of felsic rocks, i.e., the Hirnantian glacial sediments were derived from

622 rocks having granodioritic to granitic composition (see Rieu et al., 2007; Passchier and
D
TE

623 Erukanure, 2010; Nagarajan et al., 2015; Rashid and Ganai, 2015). This conclusion is

624 reinforced by the fact that chemical weathering during glacial erosion and transport is
P

625 minimal (van de Kamp, 2010) so that CIA values largely reflect the source rocks (Nesbitt and
CE

626 Young, 1996; Bahlburg and Dobrzinski, 2011). Gaschnig et al. (2014) found out that the CIA
AC

627 values of Precambrian and Palaeozoic glacial diamictites are generally higher than those of

628 fresh igneous rocks and the pervasive depletion in Sr relative to the average upper continental

629 crust demonstrates variable rate of chemical weathering of the source terrains. These glacial

630 diamictites generally show a weathering signature inherited from their provenance because

631 there has been very limited syn- or post-depositional chemical weathering (Li et al., 2016).

632 The extrabasinal components of the Hirnantian diamictites are largely dominated by

633 detrital quartz and the bulk of associated rock fragments (excluding mudstones derived from

634 the local basement) are also markedly quartz-rich. This mineralogically mature composition is

635 further emphasized by the ultrastable heavy mineral assemblage with prevailing zircon. The
ACCEPTED MANUSCRIPT
27

636 paucity of feldspar grains and abundance of detrital quartz in association with the variable but

637 generally high amount of clay matrix suggest terrigenous supply from a source area that has

638 been subjected to intensive chemical weathering (see Suttner et al., 1981; Johnson, 1993;

T
639 Nesbitt et al., 1996; van de Kamp, 2010). Although the quartz was evidently derived from

IP
640 diverse parent rocks the pronounced dominance of monocrystalline grains with uniform to

R
641 slightly undulose extinction implies a felsic plutonic source, i.e., granitoids (Basu et al.,

SC
642 1975). These conclusions about palaeoweathering and provenance based on micropetrography

643 are consistent with the above interpretation of the CIA values. However, the compositional

644
NU
maturity of the diamictites is accompanied by some textural maturity as evidenced by the
MA
645 good roundness and medium to high degree of sphericity of many detrital grains. This may

646 indicate recycling of older sediments or sedimentary rocks because the glaciomarine

647 depositional environment precludes strong reworking as an alternative explanation.


D
TE

648

649 8. Hirnantian glaciomarine record of the Balkan Terrane – a discussion


P

650
CE

651 8.1. Regional correlation, palaeogeographic position and terrane affinity


AC

652

653 An attempted correlation between the Balkan Terrane and other regions with Hirnantian

654 glaciomarine record in Europe, Africa and Asia (Fig. 12A) reveals close similarities to

655 occurrences of ice-distal deposits in France (Doré, 1981; Robardet and Doré, 1988; Barca et

656 al., 1996; Piçarra et al., 2002), Germany (Steiner and Falk, 1981; Katzung, 1999; Ziegler and

657 Wimmenauer, 2001), Italy (Leone et al., 2002; Oggiano and Mameli, 2006), Austria

658 (Schönlaub et al., 2011), and the Czech Republic (Brenchley and Štorch, 1989; Štorch, 1990).

659 The most prominent feature is the presence of fine-grained diamictites (i.e., having muddy to

660 sandy groundmass) with outsized clasts despite variations in the clast size, morphology,
ACCEPTED MANUSCRIPT
28

661 abundance and provenance described from those areas. Other common characteristics include

662 rock colour, bedding patterns, sedimentary structures, quartz enrichment, lack of fossils, and

663 lithology of the postglacial cover. The glaciomarine successions have various total

T
664 thicknesses but constantly conformable contacts with underlying and overlying marine strata.

IP
665 A close analogue of the Sirman Fm occurs in the Kučaj Unit of the East Serbian Carpatho–

R
666 Balkanides, i.e., Serbian part of the Balkan Terrane, where similar characteristics (thickness,

SC
667 colour, bedding, texture, composition) of a Hirnantian silicilastic unit and its relations with

668 the units below and above led Krstič et al. (2005) to infer deposition in a glaciomarine

669
NU
environment. In contrast to the presented evidence from other peri-Gondwana terranes with
MA
670 exposed glaciomarine strata, i.e., Morocco (Le Heron et al., 2008), Portugal (Brenchley et al.,

671 1991), Spain (Álvaro and Van Vliet-Lanoë, 2009), France (Doré, 1981), Czech Republic

672 (Štorch, 2006), Turkey (Monod et al., 2003) and Iran (Ghavidel-syooki et al., 2011), the
D
TE

673 acquired data from the Balkan Terrane allow recognition of only one major cycle of ice sheet

674 advance and retreat which produced a thin deglacial sequence. Most studies have
P

675 demonstrated that the Hirnantian included two main phases of ice advance and retreat with the
CE

676 second phase in the late Hirnantian marking the glacial maximum (see Melchin et al., 2013;
AC

677 Ghienne et al., 2014, and references therein). The late Hirnantian age (persculptus Biozone)

678 of the Sirman Fm implies that the recorded single cycle of ice advance–retreat corresponds to

679 the more extensive second phase of Hirnantian glaciation.

680 The new sedimentological results and the regional correlation unequivocally

681 demonstrate that the Hirnantian rocks of western Bulgaria were deposited on the non-

682 glaciated shelf of the North Gondwana platform. This conclusion is consistent with available

683 palaeobiogeographic data from the Ordovician System of the Balkan Terrane, for example,

684 the Middle Ordovician benthic faunas (brachiopods and trilobites) and acritarchs that are

685 closely related to peri-Gondwana provinces (Kalvacheva, 1990; Gutierrez-Marco et al., 2002,
ACCEPTED MANUSCRIPT
29

686 2003). Although the peri-Gondwana origin of three Palaeozoic terranes in Bulgaria was

687 suggested long ago by Yanev (1993), the Balkan Terrane has been rarely shown since then on

688 global or regional palaeogeographic maps, and moreover, these refer mainly to post-

T
689 Ordovician time slices of the Palaeozoic (Yanev et al., 2006; Oczlon et al., 2007; Kalvoda and

IP
690 Bábek, 2010; Sachanski et al., 2010). Only recently, Sachanski et al. (2015) used the

R
691 Hirnantian reconstruction of North Gondwana after Scotese (2014) to illustrate the

SC
692 palaeogeographic position of Silurian–Lower Devonian sequences exposed in Africa, Europe

693 and Turkey, including the Balkan Terrane. The same global palaeomap can be applied in this

694
NU
study to show the inferred location of the Balkan Terrane in the latest Ordovician (Fig. 12B)
MA
695 with juxtaposition of the other peri-Gondwana terranes where dominantly glaciomarine

696 sedimentation occurred during the Hirnantian (based on data compiled from Monod et al.,

697 2003; Yanev et al., 2006; Linnemann et al., 2011; Torsvik and Cocks, 2011; Melchin et al.,
D
TE

698 2013; Stampfli et al., 2013; Sachanski et al., 2015, and references therein).

699 Several peri-Gondwana terranes having different present-day locations in Europe were
P

700 situated along the north periphery of Gondwana, i.e., on the southern margin of the Rheic
CE

701 Ocean, during the early Palaeozoic. Some of them (Armorica s. str., Iberia, Bohemia, Saxo-
AC

702 Thuringia) have been referred to as ‘Armorican Terrane Assemblage’ (Tait et al., 1997) in the

703 literature, although the number of terranes included varies according to different authors (e.g.,

704 Franke, 2000; Yanev et al., 2006; Stampfli et al., 2013). Linnemann et al. (2000) noted that

705 this term should be used for the group of peri-Gondwana terranes showing important common

706 features, i.e., ‘Armorican affinities’, such as the occurrence of Cadomian (Pan-African)

707 basement, peri-Gondwanan Cambro-Ordovician fauna, Ordovician oolitic ironstones,

708 Hirnantian glaciomarine deposits, and absence of a ‘Caledonian unconformity’. Following

709 these considerations, Yanev et al. (2005, 2006) and Winchester et al. (2006) assigned the

710 Balkan Terrane to the Armorican Terrane Assemblage on the basis of stratigraphic,
ACCEPTED MANUSCRIPT
30

711 palaeobiogeographic and palaeoclimatologic characteristics including (besides those

712 mentioned above) the presence of late Neoproterozoic oceanic lithosphere and Cambrian–

713 early Ordovician island-arc magmatism (von Quadt et al. 1998; Savov et al., 2001; Carrigan

T
714 et al., 2003), occurrence of Ordovician oolitic ironstones (Yanev, 2000; Gutierrez-Marco et

IP
715 al., 2003), and Emsian chitinozoans showing Gondwana affinities (Yanev et al., 2005, 2006).

R
716 In the global reconstruction for the Hirnantian proposed by Stampfli et al. (2013) the Balkan

SC
717 Terrane can be integrated as part of the Armorica sub-terrane of the Galatian superterrane.

718 The Armorican Terrane Assemblage (or Armorica sub-terrane, Armorica s.l.) became

719
NU
detached from Gondwana in the Early–Middle Devonian along with closing of the Rheic
MA
720 Ocean and opening of the Palaeotethys Ocean (Robardet, 2003; Nance et al., 2010; Torsvik

721 and Cocks, 2011; Stampfli et al., 2013).

722
D
TE

723 8.2. Source of siliciclastics

724
P

725 The presence of mudstone extraclasts in the glaciomarine deposits of the Balkan
CE

726 Terrane reflects erosional processes which probably affected the locally emergent pre-
AC

727 Hirnantian basement as a result of the significant glacio-eustatic sea-level fall (see Brenchley

728 et al., 2006; Loi et al., 2010; Ghienne, 2011; Kröger et al., 2015). However, the bulk of

729 extrabasinal components clearly reveals a mineralogically mature composition of the

730 diamictites, and hence implies a specific major sediment source. This compositional maturity

731 is mainly indicated by the large predominance of detrital quartz but also by the ultrastable

732 heavy mineral assemblage, presence of quartz-rich lithics, scarcity of feldspar grains, very

733 low amount of unstable rock fragments (excluding those derived from basement lithologies),

734 and abundant clay matrix. The provenance may have been related in general to chemically

735 weathered felsic plutonic rocks which is confirmed by both the CIA values of analyzed bulk
ACCEPTED MANUSCRIPT
31

736 samples from all lithofacies and the optical characteristics of detrital quartz grains in the

737 diamictites. However, some textural maturity of the diamictites is evidenced by the roundness

738 and sphericity of many detrital grains, and therefore suggests possible recycling of older

T
739 sediments or sedimentary rocks.

IP
740 Extreme chemical weathering and erosion on the uplifted landscape of North Gondwana

R
741 took place from the latest Neoproterozoic to the pre-glacial Ordovician forming a peneplain

SC
742 and producing large volumes of quartz-rich sands (Avigad et al., 2003, 2005). These

743 sediments were derived mainly from the Neoproterozoic Pan-African continental basement

744
NU
(including abundant granitoids) and deposited as a thick sheet of first-cycle sediments across
MA
745 a vast area extending from the present-day northwestern Africa to the Arabian Peninsula

746 (Noblet and Lefort, 1990; Burke et al., 2003; Avigad et al., 2005). Meanwhile, other studies

747 based on detrital zircon geochronologies suggested contribution from more remote Gondwana
D
TE

748 source areas comprising older Precambrian crust (Avigad et al., 2003; Kolodner et al., 2006;

749 Squire et al., 2006; Linnemann et al., 2011; Morag et al., 2011; Altumi et al., 2013; Meinhold
P

750 et al., 2013; Rösel et al., 2014). Dispersal of the eroded sandy material on the margins of
CE

751 North Gondwana was favoured by thermal subsidence in the aftermath of a late
AC

752 Neoproterozoic (post-orogenic) rifting that provided accommodation for the thick blanket of

753 fluvial and marine quartz-rich sands (Burke et al., 2003). Transport during Cambrian–

754 Ordovician times occurred via a continent-wide braided river system with general south-to-

755 north direction throughout North Africa and Arabia (Beuf et al., 1971; Avigad et al., 2005;

756 Kolodner et al., 2006), and further northwards in peri-Gondwana areas (Noblet and Lefort,

757 1990; Linnemann and Romer, 2002; Avigad et al., 2012; Shaw et al., 2014). The siliciclastic

758 Cambrian–Ordovician sequences are dominated by mineralogically and often texturally

759 mature sandstones (Avigad et al., 2005; Kolodner et al., 2006; Sabaou et al., 2009;

760 Linnemann et al., 2011; Bassis et al., 2016). This maturity resulted not only from the intensive
ACCEPTED MANUSCRIPT
32

761 chemical weathering but also from the long-distance fluvial transport (Morag et al., 2011;

762 Shaw et al., 2014) and coastal reworking where deposition took place in shallow marine

763 environment (Sabaou et al., 2009). Sedimentation during the Ordovician became dominated

T
764 by quartz arenites (e.g., Avigad et al., 2005; Kolodner et al., 2006; Morag et al., 2011; Shaw

IP
765 et al., 2014) suggesting recycled sediment sources (Noblet and Lefort, 1990; Linnemann and

R
766 Romer, 2002; Sabaou et al., 2009).

SC
767 The combined analysis of micropetrographical and geochemical data obtained from

768 glaciomarine rocks of the Balkan Terrane implies that the major source of siliciclastic

769
NU
material was associated with quartz-rich sands deposited across the North Gondwana platform
MA
770 in Cambrian and pre-glacial Ordovician times. During the Hirnantian glaciation the mature

771 sandy detritus was incorporated in the expanding ice-sheet and then released from the melting

772 ice on the broad shelf area of the platform (Huuse et al., 2012). This conclusion is supported
D
TE

773 by available data for the main palaeo-ice flow orientations which indicate that the glaciogenic

774 material was transported by continental glaciers and floating ice in northwest to northeast
P

775 directions (Beuf et al., 1971; Monod et al., 2003; Moreau et al., 2005; Ghienne et al., 2007;
CE

776 Le Heron and Craig, 2008; Denis et al., 2010; Le Heron and Howard, 2010). The redeposition
AC

777 of mature siliciclastics derived from Cambrian-Ordovician sandstones in Hirnantian

778 sediments was particularly suggested for glaciogenic successions in Lybia (Le Heron and

779 Howard, 2010) and Saudi Arabia (Bassis et al., 2016). Furthermore, a literature review reveals

780 that significant quartz enrichment of Hirnantian glaciomarine diamictites was reported in a

781 number of studies of peri-Gondwanan terranes in Portugal (Brenchley et al., 1991; Couto et

782 al., 2013), Spain (Fortuin, 1984; Álvaro and Van Vliet-Lanoë, 2009), France (Robardet and

783 Doré, 1988), Sardinia (Leone et al., 2002), Czech Republic (Brenchley and Štorch, 1989),

784 Turkey (Ghienne et al., 2010), and Iran (Ghavidel-syooki et al., 2011). Published

785 micropetrographic data demonstrate that the morphology of detrital quartz grains varies from
ACCEPTED MANUSCRIPT
33

786 angular (e.g., Fortuin, 1984) to well rounded (e.g., Brenchley and Štorch, 1989). Therefore, a

787 major source of mature siliciclastics for those ice-proximal to ice-distal environments must

788 have been also related to the Cambrian-Ordovician quartz-rich sands of North Gondwana

T
789 although the patterns of sedimentary recycling and glacial/deglacial reworking have

IP
790 apparently varied.

R
791 The calculated CIA values for the studied glaciomarine deposits closely correspond to

SC
792 the composition of average shale thus indicating moderate degree of weathering in the source

793 area. However, assuming that the major source of siliciclastics was associated with

794
NU
sedimentary recycling of mature Cambro-Ordovician sands from North Gondwana is at
MA
795 variance with the extreme chemical weathering conditions which prevailed across that vast

796 area from the latest Neoproterozoic to the pre-glacial Ordovician and produced first-cycle

797 quartz arenites with abundant kaolinite (i.e., Avigad et al., 2005). This discrepancy can be
D
TE

798 explained by some contribution of terrigenous material derived from less weathered sources

799 (see Romer and Hahne, 2010) and/or erosion of the local pre-glacial basement that generated
P

800 mudstone extraclasts in the diamictites.


CE

801
AC

802 9. Conclusions

803

804 Hirnantian siliciclastic rocks exposed in the western Balkan Terrane comprise four

805 lithofacies having distinct morphological and microtextural characterisitics. The diamictites

806 and mudstones provide clear evidence for glaciomarine sedimentation in an ice-intermediate

807 to ice-distal, low-energy shelf envrionment. Deposition resulted from suspension settling from

808 meltwater plumes, rain-out of ice-rafted debris, remobilization by sediment gravity flows, and

809 probably episodic bottom currents. The glaciomarine sediments accumulated mainly during

810 retreat of a distant grounded or floating ice sheet and coeval sea-level rise in the late
ACCEPTED MANUSCRIPT
34

811 Hirnantian. A single deglaciation sequence was produced, reflecting a gradually reduced

812 sedimentation rate, waning of mass-flow processes and sediment supplied through ice-rafting,

813 and increasing contribution from suspension fallout. This glaciomarine succession is

T
814 relatively thin and laterally persistent thus representing an important stratigraphic marker in

IP
815 the Lower Palaeozoic sedimentary record of the Balkan Terrane. It also provides valuable

R
816 support for the ‘Armorican affinity’ and palaeogeographic position of this Palaeozoic terrane

SC
817 during the late Ordovician. The sedimentological analysis and correlation with Hirnantian

818 glaciogenic strata from other regions conclude that deposition occurred on the non-glaciated

819
NU
shelf of the North Gondwana platform. In a compiled palaeogeographic reconstruction of the
MA
820 broad domain with glaciomarine sedimentation along the southern margin of the Rheic ocean,

821 the Balkan Terrane is placed as a missing piece of the north peri-Gondwana jigsaw puzzle,

822 thus linking numerous terranes from Northwest Africa/Europe and remote terranes in
D
TE

823 Southwest Asia (Turkey, Iran). Mineralogical maturity of the Hirnantian diamictites and

824 additional geochemical data imply that the provenance of siliciclastic material was associated
P

825 mainly with sedimentary recycling of mature sands which had been deposited in fluvial and
CE

826 marine environments across North Gondwana in Cambrian and pre-glacial Ordovician times.
AC

827 These results can be widely used for correlation with glaciomarine strata from the

828 Hirnantian record of other North Gondwana and peri-Gondwana terranes. On a broader scale,

829 the present study enables better understanding of glaciomarine sedimentation in ancient

830 environments in terms of depositional processes and their products. In particular, the

831 investigated deposits represent a prominent example of deglaciation sequence formed entirely

832 in ice-intermediate to ice-distal settings. Therefore, they provide a solid basis for comparison

833 with similar glaciogenic successions from the geological record suggesting reliable criteria for

834 identification of diverse glaciomarine deposits especially as regards micropetrography.

835 Furthermore, this study demonstrates that combined field, microtextural and geochemical data
ACCEPTED MANUSCRIPT
35

836 may be used to thoroughly analyze the provenance of siliclastic material supplied in

837 glaciomarine environments.

838

T
839 Acknowledgements

IP
840

R
841 I greatly acknowledge Assoc. Prof. Dian Vangelov (Sofia University ‘St. Kliment Ohridski’)

SC
842 who gave the original idea for this study. I am also deeply grateful to Assoc. Prof. Valeri Sachanski

843 and Dr. Stoyan Tanatsiev (Mining-Geological University ‘St. Ivan Rilski’, Sofia) for their very

NU
844 important assistance during the field work. I thank Prof. Petr Štorch (Czech Republic) and an

845 anonymous reviewer for their helpful comments and critical remarks on the manuscript.
MA
846

847 References
D

848
TE

849 Aceñolaza, F.G., Yanev, S., 2001. El Ordovícico del sector occidental de Stara Planina (Montes

850 Balcanes), Bulgaria: Icnofósiles e implicaciones paleobiogeográficas. Revista del Museo


P
CE

851 Argentino de Ciencias Naturales 3, 73–76.

852 Allen, P., Leather, J., Brasier, M.D., 2004. The Neoproterozoic Fiq glaciation and its aftermath, Huqf
AC

853 Supergroup of Oman. Basin Research 16, 507–534.

854 Altumi, M.M., Elicki, O., Linnemann, U., Hoffman, M., Sawage, A., Gärtner, A., 2013. U-Pb LA-

855 ICP-MS detrital zircon ages from the Cambrian of Al Qarqaf Arch, central-western Libya:

856 Provenance of the West Gondwanan sand sea at the dawn of the Early Palaeozoic. Journal of

857 African Earth Sciences 79, 74– 97.

858 Álvaro, J.J., Van Vliet-Lanoë, B., 2009. Late Ordovician carbonate productivity and glaciomarine

859 record under quiescent and active extensional tectonics in NE Spain. In: Bassett, M.G. (Ed.),

860 Early Palaeozoic Peri-Gondwana terranes: New Insights from Tectonics and Biogeography.

861 Geological Society of London, Special Publications 325, pp. 117–139.


ACCEPTED MANUSCRIPT
36

862 Ampaiwan, T., Hisada, K.I., Charusiri, P., 2009. Lower Permian glacially influenced deposits in

863 Phuket and adjacent islands, peninsular Thailand. Island Arc 18, 52–68.

864 Аngelov, V., Antonov, M., Gerdzhikov, S., Tanatsiev, S., Kiselinov, H., Petrov, P., Valev, V., 2010.

865 Geological map of the Republic of Bulgaria scale 1:50,000. Map sheet Litakovo. Sofia, Apis

T
IP
866 50.

867 Аngelov, V., Antonov, M., Gerdzhikov, S., Sirakov, V., Tanatsiev, S., Sachanski, V., Valev, V., 2011.

R
868 Geological map of the Republic of Bulgaria scale 1:50,000. Map sheet Svoge. Sofia, Apis 50.

SC
869 Antić, M., Peytcheva, I., von Quadt, A., Kounov, A., Trivić, B., Serafimovski, T., Tasev, G.,

NU
870 Gerdjikov, I., Wetzel, A., 2016. Pre-Alpine evolution of a segment of the North-Gondwanan

871 margin: geochronological and geochemical evidence from the central Serbo-Macedonian
MA
872 Massif. Gondwana Research 36, 523–544.

873 Armstrong, H.A., Turner, B.R., Makhlouf, I.A., Williams, M., Al Smadi, A., Abu Salah, A., 2005.

874 Origin, sequence stratigraphy and depositional environment of an Upper Ordovician


D

875
TE

(Hirnantian), peri-glacial black shale, Jordan. Palaeogeography, Palaeoclimatology,

876 Palaeoecology 220, 273–289.


P

877 Arnaud, E. 2012. The palaeoclimatic significance of deformation structures in Neoproterozoic


CE

878 successions. Sedimentary Geology 243–244, 33–56.

879 Arnaud, E., Eyles, C.H., 2006. Neoproterozoic environmental change recorded in the Port Askaig
AC

880 Formation, Scotland: climatic vs. tectonic controls. Sedimentary Geology 183, 99–124.

881 Arnaud, E., Etienne, J.L., 2011. Recognition of glacial influence in Neoproterozoic sedimentary

882 successions. In: Arnaud, E., Halverson, G.P., Shields-Zhou, G. (Eds.), The Geological Record

883 of Neoproterozoic Glaciations. Geological Society of London, Memoirs 36, pp. 39–50.

884 Avigad, D., Kolodner, K., McWilliams, M., Persing, H., Weissbrod, T., 2003. Origin of northern

885 Gondwana Cambrian sandstone revealed by detrital zircon SHRIMP dating. Geology 31, 227–

886 230.

887 Avigad, D., Sandler, A., Kolodner, K., Stern, R.J., McWilliams, M., Miller, N., Beyth, M., 2005. Mass

888 production of Cambro-Ordovician quartz-rich sandstone as a consequence of chemical


ACCEPTED MANUSCRIPT
37

889 weathering of Pan-African terranes: environmental implications. Earth and Planetary Science

890 Letters 240, 818–826.

891 Avigad, D., Gerdes, A., Morag, N., Bechstädt, T., 2012. Coupled U–Pb–Hf of detrital zircons of

892 Cambrian sandstones from Morocco and Sardinia: implications for provenance and

T
IP
893 Precambrian crustal evolution of North Africa. Gondwana Research 21, 690–703.

894 Bahlburg, H., Dobrzinski, N., 2011. A review of the Chemical Index of Alteration (CIA) and its

R
895 application to the study of Neoproterozoic glacial deposits and climate transitions. In: Arnaud,

SC
896 E., Halverson, G.P., Shields–Zhou, G. (Eds.), The Geological Record of Neoproterozoic

NU
897 Glaciations. Geological Society of London, Memoirs, 36, pp. 81–92.

898 Barca, S., Durand-Delga, M., Rossi, P., Štorch, P., 1996. Les micaschistes panafricains de Corse et
MA
899 leur couverture paléozoïque: leur interprétation au sein de l'orogène varisque sud-européen.

900 Comptes Rendus de l'Académie des Sciences, Série 2. Sciences de la terre et des planètes 322,

901 981–989.
D

902
TE

Bassis, A., Hinderer, M., Meinhold, G., 2016. New insights into the provenance of Saudi Arabian

903 Palaeozoic sandstones from heavy mineral analysis and single-grain geochemistry.
P

904 Sedimentary Geology 333, 100–114.


CE

905 Basu, A., Young, S., Suttner, L., James, W., Mack, G., 1975. Re-evaluation of the use of undulatory

906 extinction and crystallinity in detrital quartz for provenance interpretation. Journal of
AC

907 Sedimentary Petrology 45, 873–882.

908 Bennett, M.R., Doyle, P., Mather, A.E., 1996. Dropstones: their origins and significance.

909 Palaeogeography, Palaeoclimatology, Palaeoecology 121, 331–339.

910 Bernárdez, E., Gutiérrez-Marco, J.C., Hacar, M., 2006. Sedimentos glaciomarinos del Ordovícico

911 terminal en la Zona Cantábrica (NO de España). Geogaceta 40, 239–242.

912 Beuf, S., Biju-Duval, B., de Charpal, O., Rognon, P., Gariel, O., Bennacef, A., 1971. Les grés du

913 Paléozoïque inférieur au Sahara. Sédimentation et discontinuités; évolution structurale d'un

914 craton. Publication Institut Français du Pétrole 18, 1–464.

915 Bonev, N., Ovtcharova-Schaltegger, M., Marchev, P., Moritz, R., Ulianov, A., 2013. Peri-Gondwanan

916 Ordovician crustal fragments in the high-grade basement of the Eastern Rhodope Massif,
ACCEPTED MANUSCRIPT
38

917 Bulgaria: evidence from U-Pb LA-ICP-MS zircon geochronology and geochemistry.

918 Geodinamica Acta 26, 207–229.

919 Boulton, G.S., 1990. Sedimentary and sea level changes during glacial cycles and their control on

920 glacimarine facies architecture. In: Dowdeswell, J.A., Scourse, J.D. (Eds.), Glacimarine

T
IP
921 Environments: Processes and sediments. Geological Society of London, Special Publications

922 53, pp. 15–52.

R
923 Brenchley, P.J., Štorch, P., 1989. Environmental changes in the Hirnantian (upper Ordovician) of the

SC
924 Prague Basin, Czechoslovakia. Geological Journal 24, 165–181.

NU
925 Brenchley, P.J., Romano, M., Young, T.P., Štorch, P., 1991. Hirnantian glaciomarine diamictites –

926 evidence for the spread of glaciation and its effect on Upper Ordovician faunas. In: Barnes,
MA
927 C.R., Williams, S.H. (Eds.), Advances in Ordovician geology. Geological Survey Canada Paper

928 90–9, pp. 325−336.

929 Brenchley, P.J., Marshall, J.D., Harper, D.A., Buttler, C.J., Underwood, C.J., 2006. A late Ordovician
D

930
TE

(Hirnantian) karstic surface in a submarine channel, recording glacio-eustatic sea-level

931 changes: Meifod, central Wales. Geological Journal 41, 1–22.


P

932 Burke, K., MacGregor, D.S., Cameron, N.R., 2003. Africa’s petroleum systems: four tectonic ‘Aces’
CE

933 in the past 600 million years. In: Arthur, T.J., MacGregor, D.S., Cameron, N.R. (Eds.),

934 Petroleum geology of Africa: new themes and developing technologies. Geological Society of
AC

935 London, Special Publications 207, pp. 21–60.

936 Busfield, M.E., Le Heron, D. P., 2013. Glacitectonic deformation in the Chuos Formation of northern

937 Namibia: implications for Neoproterozoic ice dynamics. Proceedings of the Geologists’

938 Association 124, 778–789.

939 Carr, S.J., 2001. Micromorphological criteria for distinguishing subglacial and glacimarine sediments:

940 evidence from a contemporary tidewater glacier, Spitsbergen. Quaternary International 86, 71–

941 79.

942 Carr, S.J., Haflidason, H., Sejrup, H.P., 2000. Micromorphological evidence supporting late

943 Weichselian glaciation of the northern North Sea. Boreas 29, 315–328.
ACCEPTED MANUSCRIPT
39

944 Carrigan, C.W., Mukasa, S., Haydoutov, I., Kolcheva, K., 2003. Ion microprobe U-Pb zircon ages of

945 pre-Alpine rocks in the Balkan, Sredna gora and the Rhodope terranes of Bulgaria: Constraints

946 on Neoproterozoic and Variscan tectonic evolution. Journal of Czech Geological Society 48,

947 32–33.

T
IP
948 Chatalov, A. 2014. Development of strain fringes in sedimentary rocks: Evidence for deformation of

949 Upper Ordovician glacial diamictites in the western Srednogorie Zone. Geologica Balcanica

R
950 43, 51–62.

SC
951 Chumakov, N.M., 1985. Glacial events of the past and their geological significance. Palaeogeography,

NU
952 Palaeoclimatology, Palaeoecology 51, 319–346.

953 Condie, K.C., Des Marais D.J., Abbott, D., 2001. Precambrian superplumes and supercontinents: A
MA
954 record in black shales, carbon isotopes, and paleoclimates? Precambrian Research 106, 239–

955 260.

956 Couto, H., Knight, J., Lourenço, A., 2013. Late Ordovician ice-marginal processes and sea-level
D

957
TE

change from the north Gondwana platform: evidence from the Valongo Anticline (northern

958 Portugal). Palaeogeography, Palaeoclimatology, Palaeoecology 375, 1–15.


P

959 Cowan, E.A., Christoffersen, P., Powell, R.D., 2012. Sedimentological signature of a deformable bed
CE

960 preserved beneath an ice stream in a late Pleistocene glacial sequence, Ross Sea, Antarctica.

961 Journal of Sedimentary Research 82, 270–282.


AC

962 Crowell, J.C., 1999. Pre-Mesozoic Ice Ages: Their Bearing on Understanding the Climate System.

963 Geological Society of America, Memoir 192, 106 pp.

964 Denis, M., Buoncristiani, J.-F., Konaté, M., Ghienne, J.-F., Guiraud, M., 2007. Hirnantian glacial and

965 deglacial record in SW Djado Basin (NE Niger). Geodinamica Acta 20, 177–195.

966 Denis, M., Guiraud, M., Konate, M., Buoncristiani, J.F., 2010. Subglacial deformation and water-

967 pressure cycles as a key for understanding ice stream dynamics: Evidence from the Late

968 Ordovician succession of the Djado Basin (Niger). International Journal of Earth Sicneces 99,

969 1399–1425.

970 Deschamps, R., Eschard, R., Roussé, S., 2013. Architecture of Late Ordovician glacial valleys in the

971 Tassili N'Ajjer area (Algeria). Sedimentary Geology 289, 124–147.


ACCEPTED MANUSCRIPT
40

972 Deynoux, M., Miller, J.M.G., Domack, E.W., Eyles, N., Fairchild, I.J., Young, G.M. (Eds.), 1994.

973 Earth's Glacial Record. Cambridge University Press, Cambridge.

974 Dineen, A.A., Fraiser, M.L., Isbell, J.L., 2013. Palaeoecology and sedimentology of Carboniferous

975 glacial and post-glacial successions in the Paganzo and Río Blanco basins of northwestern

T
Argentina. In: Gąsiewicz, A., Słowakiewicz, M. (Eds.), Palaeozoic Climate Cycles: Their

IP
976

977 Evolutionary and Sedimentological Impact. Geological Society of London, Special

R
978 Publications 376, pp. 109–140.

SC
979 Ding, H., Ma, D., Yao, C., Lin, Q., Jing, L., 2016. Implication of the chemical index of alteration as a

NU
980 paleoclimatic perturbation indicator: an example from the lower Neoproterozoic strata of Aksu,

981 Xinjiang, NW China. Geosciences Journal 20, 13–26.


MA
982 Dobrzinski, N., Bahlburg, H., 2007. Sedimentology and environmental significance of the Cryogenian

983 successions of the Yangtze Platform, South China block. Palaeogeography, Palaeoclimatology,

984 Palaeoecology 254, 100–122.


D

985
TE

Dobrzinski, N., Bahlburg, H., Strauss, H., Zhang, Q.R., 2004. Geochemical climate proxies applied to

986 the Neoproterozoic glacial succession on theYangtze Platform, South China. In: Jenkins, G.S.,
P

987 McMenamin, M.A.S., McKay, C.P., Sohl, L.E. (Eds.), The Extreme Proterozoic: Geology,
CE

988 Geochemistry and Climate. American Geophysical Union Monograph Series 146, pp. 13–32.

989 Doré, F., 1981. The late Ordovician tillite in Normandy (Armoricain Massif). In: Hambrey, M.J.,
AC

990 Harland, W.B. (Eds.), Earth's Pre-Pleistocene Glacial Record. Cambridge University Press,

991 Cambridge, pp. 582–585.

992 Dowdeswell, J.A., Elverhøi, A., Spielhagen, R., 1998. Glacimarine sedimentary processes and facies

993 on the Polar North Atlantic margins. Quaternary Science Reviews 17, 243–272.

994 Dowdeswell, J.A., Whittington, R.J., Jennings, A.E., Andrews, J.T., Mackensen, A., Marienfeld, P.,

995 2000. An origin for laminated glacimarine sediments through sea-ice build-up and suppressed

996 iceberg rafting. Sedimentology 47, 557–576.

997 Elverhøi, A., Henrich, R., 2002. Past glaciomarine environments. In: Menzies, J. (Ed.), Modern and

998 Past Glacial Environments. Revised student edition. Butterworth-Heinemann, Oxford, pp. 391–

999 415.
ACCEPTED MANUSCRIPT
41

1000 Eyles, N., 1993. Earth’s glacial record and its tectonic setting. Earth-Science Reviews 35, 1–248.

1001 Eyles, N., 2008. Glacio-epochs and the supercontinent cycle after ~3.0 Ga: tectonic boundary

1002 conditions for glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 258, 89–129.

1003 Eyles, N., Eyles, C.H., 1992. Glacial depositional systems. In: Walker, R.G., James, M.P. (Eds.),

T
IP
1004 Facies Models: Response to Sea Level Change. Geological Association of Canada, pp. 73–100.

1005 Eyles, N., Januszczak, N., 2004. ‘Zipper-rift’: a tectonic model for Neoproterozoic glaciations during

R
1006 the breakup of Rodinia after 750 Ma. Earth-Science Reviews 65, 1–73.

SC
1007 Eyles, C.H., Eyles, N., 2010. Glacial deposits. In: James, N.P., Dalrymple, R.W.

NU
1008 (Eds.), Facies Models 4. Geological Association of Canada, pp. 73–104.

1009 Eyles, C.H., Eyles, N., Miall, A.D., 1985. Models of glaciomarine sedimentation and their application
MA
1010 to the interpretation of ancient glacial sequences. Palaeogeography, Palaeoclimatology,

1011 Palaeoecology 51, 15–84.

1012 Fedo, C.M., Young, G.M., Nesbitt, G.M., 1997. Paleoclimatic control on the composition of the
D

1013
TE

Paleoproterozoic Serpent Formation, Huronian Supergroup, Canada: a greenhouse to icehouse

1014 transition. Precambrian Research 86, 201–223.


P

1015 Fielding, C.R., Blackstone, B.A., Frank, T.D., Gui, Z., 2012. Reservoir potential of sands formed in
CE

1016 glaciomarine environments: an analogue study based on Cenozoic examples from McMurdo

1017 Sound, Antarctica. In: Huuse, M., Le Heron, D.P., Dixon, R., Redfern, J., Moscariello, A.,
AC

1018 Craig, J. (Eds.), Glaciogenic Reservoirs and Hydrocarbon Systems. Geological Society of

1019 London, Special Publications 368, pp. 211–228.

1020 Fortuin, A.R., 1984. Late Ordovician glaciomarine deposits (Orea Shale) in the Sierra de Albarracin,

1021 Spain. Palaeogeography, Palaeoclimatology, Palaeoecology 48, 245–261.

1022 Franke, W., 2000. The mid-European segment of the Variscides: Tectono-stratigraphic units, terrane

1023 boundaries and plate tectonic evolution. In: Franke, W., Haak, V., Oncken, O., Tanner, D.

1024 (Eds.), Orogenic Processes: Quantification and Modelling in the Variscan Belt. Geological

1025 Society of London, Special Publications 179, pp. 35–62.


ACCEPTED MANUSCRIPT
42

1026 Gaschnig, R.M., Rudnick, R.L., McDonough, W.F., Kaufman, A.J., Hu, Z.C., Gao, S., 2014. Onset of

1027 oxidative weathering of continents recorded in the geochemistry of ancient glacial diamictites.

1028 Earth and Planetary Science Letters 408, 87–99.

1029 Ghavidel-syooki, M., Javier Álvaro, J., Popov, L., Ghobadi Pour, M., Ehsani, M.H., Suyarkova, A.,

T
IP
1030 2011. Stratigraphic evidence for the Hirnantian (latest Ordovician) glaciation in the Zagros

1031 Mountains, Iran. Palaeogeography, Palaeoclimatology, Palaeoecology 307, 1–16.

R
1032 Ghienne, J.-F., 2003. Late Ordovician sedimentary environments, glacial cycles, and post-glacial

SC
1033 transgression in the Taoudeni Basin, West Africa. Palaeogeography, Palaeoclimatology,

NU
1034 Palaeoecology 189, 117–145.

1035 Ghienne, J.-F., 2011. The Late Ordovician glacial record: State of the art. In: Gutiérrez-Marco, J.C.,
MA
1036 Rábano, I., García-Bellido, D. (Eds.), Ordovician of the World. Madrid. Publicaciones del

1037 Instituto Geológico y Minero de España 14, pp. 13–19.

1038 Ghienne, J.-F., Le Heron, D., Moreau, J., Denis, M., Deynoux, M., 2007. The Late Ordovician glacial
D

1039
TE

sedimentary system of the North Gondwana platform. In: Hambrey, M.J., Christoffersen, P.,

1040 Glasser, N.F., Hubbard, B. (Eds.), Glacial Sedimentary Processes and Products. International
P

1041 Association of Sedimentologists, Specical Publication 39, pp. 295–319.


CE

1042 Ghienne, J.-F., Monod, O., Kozlu, H., Dean, W.T., 2010. Cambrian–Ordovician depositional

1043 sequences in the Middle East: a perspective from Turkey. Earth-Science Reviews 101, 101–
AC

1044 146.

1045 Ghienne, J.-F., Moreau, J., Degermann, L., Rubino, J.-L., 2013. Lower Palaeozoic unconformities in

1046 an intracratonic platform setting: glacial erosion versus tectonics in the eastern Murzuq Basin

1047 (southern Libya). International Journal of Earth Sciences 102, 455–482.

1048 Ghienne, J.-F., Desrochers, A., Vandenbroucke, T.R.A., Achab, A., Asselin, E., Dabard, M.-P., Farley,

1049 C., Loi, A., Paris, F., Wickson, S., Veizer, J., 2014. A Cenozoic-style scenario for the end-

1050 Ordovician glaciation. Nature Communications 5, 1–9. doi: 10.1038/ncomms5485

1051 Gilbert, R., 1990. Rafting in glacimarine environments. In: Dowdeswell, J.A., Scourse, J.D. (Eds.),

1052 Glacimarine Environments: Processes and Sediments. Geological Society of London, Special

1053 Publications 53, pp. 105–121.


ACCEPTED MANUSCRIPT
43

1054 Girard, F., Ghienne, J.-F., Du-Bernard, X., Rubino, J.-L., 2015. Sedimentary imprints of former ice-

1055 sheet margins: insights from an end-Ordovician archive (SW Libya). Earth-Science

1056 Reviews 148, 259–289.

1057 Goldschmidt, P.M., Pfirman, S.L., Wollenburg, I., Henrich, R., 1992. Origin of sediment pellets from

T
IP
1058 the Arctic seafloor: Sea ice or icebergs? Deep Sea Research Part A. Oceanographic Research

1059 Papers 39 (2, Part 1), 539–565.

R
1060 Gutiérez-Marco, J.C., Yanev, S., Sachanski, V., Rabano, I., Lakova, I., 2002. New findings of

SC
1061 trilobites and graptolites in the Ordovician of Bulgaria. Review of the Bulgarian Geological

NU
1062 Society 63, 51–58 (in Bulgarian with an English abstract)

1063 Gutiérez-Marco, J.C., Yanev, S.N., Sachanski, V.V., Rabano, I., Lakova, I., San Jose Lancha, M.A.,
MA
1064 Diez Martinez, E., Boncheva, I., Sarmiento, G.N., 2003. New biostratigraphical data from the

1065 Ordovician of Bulgaria. INSUGEO, Serie Correlación Geológica 17, 79–85.

1066 Gutiérrez-Marco, J.C., Ghienne, J.-F., Bernárdez, E., Hacar, M.P., 2010. Did the Late Ordovician
D

1067
TE

Africa ice sheet reach Europe? Geology 38, 279–282.

1068 Haberfelner, E., Bončev, E., 1934. Der erste Nachweis von Ordovicium in Bulgarien:
P

1069 Didymograptenschiefer mit Trilobiten im Zerie-Massiv. Geologica Balcanica 1, 28–33.


CE

1070 Halverson, G.P., Maloof, A.C., Hoffman, P.F., 2004. The Marinoan glaciation (Neoproterozoic) in

1071 northeast Svalbard. Basin Research 16, 297–324.


AC

1072 Hambrey, M. J., 1994. Glacial environments. UCL Press Ltd., London.

1073 Hambrey, M.J., Harland, W.B. (Eds.), 1981. Earth's Pre-Pleistocene Glacial Record. Cambridge

1074 University Press, Cambridge.

1075 Hambrey, M.J., Glasser, N.F., 2003. Glacial sediments: processes, environments and facies. In:

1076 Middleton, G.V. (Ed.), Encyclopedia of Sediments and Sedimentary Rocks. Springer,

1077 Dordrecht, pp. 316–331.

1078 Harris, C., 1998. The micromorphology of paraglacial and periglacial slope deposits: a case study

1079 from Morfa Bychan, west Wales, UK. Journal of Quaternary Science 13, 73–84.

1080 Haydoutov, I., Yanev, S., 1997. The Protomoesian microcontinent of Balkan Peninsula – a peri-

1081 Gondwanaland piece. Tectonophysics 272, 303–313.


ACCEPTED MANUSCRIPT
44

1082 Henry, L.C., Isbell, J.L., Limarino, C.O., McHenry, L.J., Fraiser, M.L., 2010. Mid-Carboniferous

1083 deglaciation of the Protoprecordillera, Argentina recorded in the Agua de Jagüel palaeovalley.

1084 Palaeogeography, Palaeoclimatology, Palaeoecology 298, 112–129.

1085 Henry, L.C., Isbell, J.L., Fielding, C.R., Domack, E.W., Frank, T.D., Fraiser, M.L., 2012. Proglacial

T
IP
1086 deposition and deformation in the Upper Carboniferous to Lower Permian Wynyard Formation,

1087 Tasmania: a process analysis. Palaeogeography, Palaeoclimatology, Palaeoecology 315– 316,

R
1088 142–157.

SC
1089 Hiemstra, J., Rijsdijk, K., 2003. Observing artificially induced strain: implications for subglacial

NU
1090 deformation. Journal of Quaternary Science 18, 373–383.

1091 Himmerkus, F., Reischmann, T., Kostopoulos, D., 2009. Serbo-Macedonian revisited: a Silurian
MA
1092 basement terrane from northern Gondwana in the Internal Hellenides, Greece. Tectonophysics

1093 473, 20–35.

1094 Huuse, M., Le Heron, D.P., Dixon, R., Redfern, J., Moscariello, A., Craig, J., 2012. Glaciogenic
D

1095
TE

reservoirs and hydrocarbon systems: an introduction. In: Huuse, M., Redfern, J., Le Heron,

1096 D.P., Dixon, R., Moscariello, A., Craig, J. (Eds.), Glaciogenic Reservoirs and Hydrocarbon
P

1097 Systems. Geological Society of London, Special Publications 368, pp. 1–28.
CE

1098 Isbell, J.L., 2010. Environmental and paleogeographic implications of glaciotectonic deformation of

1099 glaciomarine deposits within Permian strata of the Metschel Tillite, southern Victoria Land,
AC

1100 Antarctica. In: López-Gamundí, O.R., Buatois, L.A. (Eds.), Late Paleozoic Glacial Events and

1101 Postglacial Transgressions in Gondwana. Geological Society of America, Special Paper 468,

1102 pp. 81–100.

1103 Isbell, J.L., Henry, L.C., Reid, C.M., Fraiser, M.L., 2013. Sedimentology and palaeoecology of

1104 lonestone-bearing mixed clastic rocks and cold-water carbonates of the Lower Permian Basal

1105 Beds at Fossil Cliffs, Maria Island, Tasmania (Australia): insight into the initial decline of the

1106 late Palaeozoic ice age. In: Gąsiewicz, A., Slowakiewicz, M. (Eds.), Palaeozoic Climate

1107 Cycles: Their Evolutionary, Sedimentological and Economic Impact. Geological Society of

1108 London, Special Publications 376, pp. 307–341.


ACCEPTED MANUSCRIPT
45

1109 Isbell, J.L., Biakov, A.S., Vedernikov, I.L., Davydov, V.I., Gulbranson, E.L., Fedorchuk, N.D., 2016.

1110 Permian diamictites in northeastern Asia: their significance concerning the bipolarity of the late

1111 Paleozoic ice age. Earth-Science Reviews 154, 279–300.

1112 Ivanov, Ž., 1988. Basic structural features of the External zones of the Western Balkanides. In:

T
Ivanov, Ž. (Ed.), Lineaments as Joint Structures of Folded Areas of Different Age and Their

IP
1113

1114 Metallogeny. Bulgarian Academy of Sciences, Sofia, pp. 49–81 (in Russian)

R
1115 Johnsson, M.J., 1993. The system controlling the composition of clastic sediments. In: Johnsson, M.J.,

SC
1116 Basu, A. (Eds.), Processes Controlling the Composition of Clastic Sediment. Geological

NU
1117 Society of America, Special Paper 284, pp. 1–19.

1118 Kalvacheva, R., 1990. Review on microfossil dating of low metamorphic rocks in Stara planina and
MA
1119 Vakarel hills. In: Nikolov, T. (Ed.), Microfossils in Bulgarian Stratigraphy, Bulgarian Academy

1120 of Sciences, Sofia, pp. 13-22 (in Russian)

1121 Kalvoda, J., Bábek, O., 2010. The margins of Laurussia in central and southeast Europe and southwest
D

1122
TE

Asia. Gondwana Research 17, 526–545.

1123 Katzung, G., 1999. Records of the Late Ordovician glaciation from Thuringia, Germany. Zeitschrift
P

1124 der Deutschen Geologischen Gesellschaft 150, 595–617.


CE

1125 Kellerhals, P., Matter, A., 2003. Facies analysis of a glaciomarine sequence, the Neoproterozoic

1126 Mirbat Sandstone Formation, Sultanate of Oman. Eclogae Geologicae Helvetiae 96, 49–70.
AC

1127 Kilfeather, A.A., Ó Cofaigh, C., Dowdeswell, J.A., van der Meer, J.J.M., Evans, D.J.A., 2010.

1128 Micromorphological characteristics of glacimarine sediments: implications for distinguishing

1129 genetic processes of massive diamicts. Geo-Marine Letters 30, 77–97.

1130 Koch, Z.J., Isbell, J.L., 2013. Processes and products of grounding-line fans from the Permian Pagoda

1131 Formation, Antarctica: insight into glacigenic conditions in polar Gondwana. Gondwana

1132 Research 24, 161–172.

1133 Kolodner, K., Avigad, D., McWilliams, M., Wooden, J.L., Weissbrod, T., Feinstein, S., 2006.

1134 Provenance of north Gondwana Cambrian–Ordovician sandstone: U–Pb SHRIMP dating of

1135 detrital zircons from Israel and Jordan. Geological Magazine 143, 367–391.
ACCEPTED MANUSCRIPT
46

1136 Kröger, B., Ebbestad, J.O.R., Lehnert, O., Ullmann, C.V., Korte, C., Frei, R., Rasmussen, C.M.Ø.,

1137 2015. Subaerial speleothems and deep karst in central Sweden linked to Hirnantian glaciations.

1138 Journal of the Geological Society 172, 349–56.

1139 Krstič, B., Maslarevič, L., Sudar, M., 2005. On the Graptolite Schists Formation (Silurian–Lower

T
IP
1140 Devonian) in the Carpatho–Balkanides of eastern Serbia. Annales Géologiques Péninsule

1141 Balkanique 66, 1–89.

R
1142 Laberg, J.S., Vorren, T.O., 2000. Flow behaviour of the submarine glacigenic debris flows on the Bear

SC
1143 Island Trough Mouth Fan, western Barents Sea. Sedimentology 47, 1105–1117.

NU
1144 Lakova, I., Sachanski, V., 2004. Cryptospores and trilite spores in oceanic graptolite-bearing

1145 sediments (Saltar Formation) across the Ordovician–Silurian boundary in the West Balkan
MA
1146 Mountains, Bulgaria. Review of the Bulgarian Geological Society 65, 151–156.

1147 Lang, J., Dixon, R.J., Le Heron, D.P., Winsemann, J., 2012. Depositional architecture and sequence

1148 stratigraphic correlation of Upper Ordovician glaciogenic deposits, Illizi Basin, Algeria. In:
D

1149
TE

Huuse, M., Redfern, J., Le Heron, D.P., Dixon, R., Moscariello, A., Craig, J. (Eds.),

1150 Glaciogenic Reservoirs and Hydrocarbon Systems. Geological Society of London, Special
P

1151 Publications 368, pp. 293–317.


CE

1152 Lea, J.M., Palmer, A., 2014. Quantification of turbate microstructures through a subglacial till:

1153 dimensions and characteristics. Boreas 43, 869–881.


AC

1154 Le Heron, D.P., Craig, J., 2008. First order reconstructions of a Late Ordovician Saharan ice sheet.

1155 Journal of the Geological Society of London 165, 19–30.

1156 Le Heron, D., Dowdeswell, J.A., 2009. Calculating ice volumes and ice flux to constrain the

1157 dimensions of a 440 Ma North African ice sheet. Journal of the Geological Society 166, 277–

1158 281.

1159 Le Heron, D.P., Howard, J., 2010. Evidence for Late Ordovician glaciation of Al Kufrah Basin, Libya.

1160 Journal of African Earth Sciences 58, 354–364.

1161 Le Heron, D.P., Ghienne, J.-F., El Houicha, M., Khoukhi, Y., Rubino, J.-L., 2007. Maximum extent of

1162 ice sheets in Morocco during the Late Ordovician glaciation. Palaeogeography,

1163 Palaeoclimatology, Palaeoecology 245, 200–226.


ACCEPTED MANUSCRIPT
47

1164 Le Heron, D.P., Khoukhi, Y., Paris, F., Ghienne, J.-F., Le Hérissé, A., 2008. Black shale, grey shale,

1165 fossils and glaciers: anatomy of the Upper Ordovician–Silurian succession in the Tazzeka

1166 Massif of eastern Morocco. Gondwana Research 14, 483–496.

1167 Le Heron, D.P., Craig, J., Etienne, J.L., 2009. Ancient glaciations and hydrocarbon accumulations in

T
IP
1168 North Africa and the Middle East. Earth-Science Reviews 93, 47–76.

1169 Le Heron, D.P., Armstrong, H.A., Wilson, C., Howard, J.P.,Gindre, L., 2010. Glaciation and

R
1170 deglaciation of the Libyan Desert: the Late Ordovician record. Sedimentary Geology 223, 100–

SC
1171 125.

NU
1172 Le Heron, D.P., Cox, G.M., Trundley, A.E., Collins, A., 2011. Two Cryogenian glacial successions

1173 compared: aspects of the Sturt and Elatina sediment records of South Australia. Precambrian
MA
1174 Research 186, 147–168.

1175 Le Heron, D.P., Busfield, M.E., Collins, A.S., 2014. Bolla Bollana Boulder Beds: a Neoproterozoic

1176 trough mouth fan in South Australia? Sedimentology 61, 978–995.


D

1177
TE

Leone, F., Ferretti, A., Hammann, W., Loi, A., Pillola, G.L., Serpagli, E., 2002. A general view on the

1178 post-Sardic Ordovician sequence from SW Sardinia. In: Cherchi, A., Corradini, C., Putzu, M.T.
P

1179 (Eds.), Sardinia Field Trip – Palaeontology and Stratigraphy. Rendiconti della Società
CE

1180 Paleontologica Italiana 1, pp. 51–68.

1181 Li, D.W., Ma, B.Q., Jiang, F.Q., Wang, P.L., 2011. Nature, genesis and provenance of silt pellets on
AC

1182 the ice surface of glacier No. 1, upper Urumqi River, Tian Shan, Northwestern China.

1183 Quaternary International 236, 107–115.

1184 Li, S., Gaschnig, R.M., Rudnick, R.L., 2016. Insights into chemical weathering of the upper

1185 continental crust from the geochemistry of ancient glacial diamictites. Geochimica

1186 Cosmochimica Acta 176, 96–117.

1187 Licht, K., 2009. Glaciomarine Sediments. In: Gornitz, V. (Ed.), Encyclopedia of Paleoclimatology and

1188 Ancient Environments. Springer, Dordrecht, pp. 395–396.

1189 Linnemann, U., Romer, R.L., 2002. The Cadomian Orogeny in Saxo-Thuringia, Germany:

1190 Geochemical and Nd–Sr–Pb isotopic characterisation of marginal basins with constraints to

1191 geotectonic setting and provenance. Tectonophysics 352, 33–64.


ACCEPTED MANUSCRIPT
48

1192 Linnemann, U., Gehmlich, M., Tichomirowa, M., Buschmann, B., Nasdala, L., Jonas, P., Lützner, H.,

1193 Bombach, K., 2000. From Cadomian subduction to early Paleozoic rifting: The evolution of

1194 Saxo-Thuringia at the margin of Gondwana in the light of single zircon geochronology and

1195 basin development (central European Variscides, Germany). In: Franke, W., Haak, V., Oncken,

T
IP
1196 O., Tanner, D. (Eds.), Orogenic Processes: Quantification and Modelling in the Variscan Belt.

1197 Geological Society of London, Special Publication 179, pp. 131–153.

R
1198 Linnemann, U., Ouzegane, K., Drareni, A., Hofmann, M., Becker, S., Gärtner, A., Sagawe, A., 2011.

SC
1199 Sands of West Gondwana: An archive of secular magmatism and plate interactions – a case

NU
1200 study from the Cambro-Ordovician section of the Tassili Ouan Ahaggar (Algerian Sahara)

1201 using U–Pb–LA-ICP-MS detrital zircon ages. Lithos 123, 188–203.


MA
1202 Loi, A., Ghienne, J.-F., Dabard, M.P., Paris, F., Botquelen, A., Christ, N., Elaouad-Debbaj, Z., Gorini,

1203 A., Vidal, M., Videt, B., Destombes, J., 2010. The Late Ordovician glacioeustatic record from a

1204 high-latitude storm-dominated shelf succession: the BouIngarf section (Anti-Atlas, Southern
D

1205
TE

Morocco). Palaeogeography, Palaeoclimatology, Palaeoecology 296, 332–358.

1206 López-Gamundí, O.R., 1991. Thin-bedded diamictites in the glaciomarine Hoyada Verde Formation
P

1207 (Carboniferous), Calingasta-Uspallata Basin, western Argentina: a discussion on the


CE

1208 emplacement conditions of subaqueous cohesive debris flows. Sedimentary Geology 73, 247–

1209 256.
AC

1210 López-Gamundí, O.R., 2010. Transgressions related to the demise of the late Paleozoic Ice Age: their

1211 sequence stratigraphic context. In: López-Gamundí, O.R., Buatois, L.A. (Eds.), Late Paleozoic

1212 Glacial Events and Postglacial Transgressions in Gondwana. Geological Society of America,

1213 Special Paper 468, pp. 1–35.

1214 Lønne, I., 1995. Sedimentary facies and depositional architecture of ice-contact glaciomarine systems.

1215 Sedimentary Geology 98, 13–43.

1216 Lüning, S., Craig, J., Loydell, D.K., Štorch, P., Fitches, B., 2000. Lower Silurian ‘hot shales’ in North

1217 Africa and Arabia: regional distribution and depositional model. Earth-Science Reviews 49,

1218 121–200.
ACCEPTED MANUSCRIPT
49

1219 Marenssi, S.A., Tripaldi, A., Limarino, C.O., Caselli, A.T., 2005. Facies and architecture of a

1220 Carboniferous grounding-line system from the Guandacol Formation, Paganzo Basin,

1221 northwestern Argentina. Gondwana Research 8, 187–202.

1222 McLennan, S.M., 1993. Weathering and global denudation. Journal of Geology 101, 295–303.

T
IP
1223 Meinhold, G., Morton, A.C., Avigad, D., 2013. New insights into peri-Gondwana paleogeography and

1224 the Gondwana super-fan system from detrital zircon U–Pb ages. Gondwana Research 23, 661–

R
1225 665.

SC
1226 Melchin, M.J., Mitchell, C.E., Holmden, C., Štorch, P., 2013. Environmental changes in the Late

NU
1227 Ordovician-Early Silurian: review and new insights from black shales and nitrogen isotopes.

1228 Geological Society of America Bulletin 125, 1635–1670.


MA
1229 Melchin, M.J., Mitchell, C.E., Naczk-Cameron, A., Fan, J.X., Loxton, J., 2011. Phylogeny and

1230 adaptive radiation of the Neograptina (Graptoloida) during the Hirnantian mass extinction and

1231 Silurian recovery. Proceedings of the Yorkshire Geological Society 58, 281–309.
D

1232
TE

Menzies, J., 2000. Micromorphological analyses of microfabrics and microstructures, indicative of

1233 deformation processes, in glacial sediments. In: Maltman, A.J., Hubbard, B., Hambrey, M.J.
P

1234 (Eds.), Deformation of Glacial Materials. Geological Society, London, Special Publications,
CE

1235 176, pp. 245–258.

1236 Menzies, J., 2002. Glacial environments — modern and past. In: Menzies, J. (Ed.), Modern and Past
AC

1237 Glacial Environments. Revised student edition. Butterworth-Heinemann, Oxford, pp. 1–13.

1238 Menzies, J., Zaniewski, K., 2003. Microstructures within modern debris flow deposits derived from

1239 Quaternary glacial diamicton – a comparative micromorphology study. Sedimentary Geology

1240 157, 31–48.

1241 Milicevic, V., 1994. Preliminary palaeomagnetic results for Ordovician of Zvonyachka Banya,

1242 Dgerchek and Zrna reka (Eastern Serbia). Proceedings Geoinstitute 29, 13–22 (in Serbian with

1243 an English abstract)

1244 Moncrieff, A.C.M., 1989. Classification of poorly sorted sedimentary rocks. Sedimentary Geology 65,

1245 191–194.
ACCEPTED MANUSCRIPT
50

1246 Monod, O., Kozlu, H., Ghienne, J.-F., Dean, W.T., Günay, Y., Le Hérissé, A., Paris, F., Robardet, M.,

1247 2003. Late Ordovician glaciation in southern Turkey. Terra Nova 15, 249–257.

1248 Morag, N., Avigad, D., Gerdes, A., Belousova, E., Harlavan, Y., 2011. Detrital zircon Hf isotopic

1249 composition indicates long-distance transport of North Gondwana Cambrian–Ordovician

T
IP
1250 sandstones. Geology 39, 955–958.

1251 Moreau, J., 2011. The late Ordovician deglaciation sequence of the SW Murzuq Basin, Libya. Basin

R
1252 Research 23, 449–477.

SC
1253 Moreau, J., Ghienne, J.-F., Le Heron, D., Rubino, J.-L., Deynoux, M., 2005. A 440Ma old ice stream

NU
1254 in North Africa. Geology 33, 753–756.

1255 Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary density flows and
MA
1256 their deposits. Sedimentology 48, 269–299.

1257 Nagarajan, R., Armstrong-Altrin, J.S., Kessler, F.L., Hidalgo-Moral, E.L., Dodge-Wan, D., Taib, N.I.,

1258 2015. Provenance and tectonic setting of Miocene siliciclastic sediments, Sibuti Formation,
D

1259
TE

northwestern Borneo. Arabian Journal of Geosciences 8, 8549–8565.

1260 Nance, R.D., Gutiérrez-Alonso, G., Keppie, J.D., Linnemann, U., Murphy, J.B., Quesada, C.,
P

1261 Strachan, R.A., Woodcock, N.H., 2010. Evolution of the Rheic Ocean. Gondwana Research 17,
CE

1262 194–222.

1263 Nelson, A.E., Smellie, J.L., Hambrey, M.J., Williams, M., Vautravers, M., Salzmann, U., McArthur,
AC

1264 J.M., Regelous, M., 2009. Neogene glacigenic debris flows on James Ross Island, northern

1265 Antarctic Peninsula, and their implications for regional climate history. Quaternary Science

1266 Reviews 28, 3138–3160.

1267 Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred from major

1268 element chemistry of lutites. Nature 199, 715–717.

1269 Nesbitt, H.W., Young, G.M., 1984. Prediction of some weathering trends of plutonic and volcanic

1270 rocks based on thermodynamic and kinetic considerations. Geochimica Cosmochimica Acta 48,

1271 1523–1534.
ACCEPTED MANUSCRIPT
51

1272 Nesbitt, H.W., Young, G.M., 1996. Petrogenesis of sediments in the absence of chemical weathering:

1273 effects of abrasion and sorting on bulk composition and mineralogy. Sedimentology 42, 341–

1274 358.

1275 Nesbitt, H.W., Young, G.M., McLennan, S.M., Keays, R.R., 1996. Effects of chemical weathering and

T
IP
1276 sorting on the petrogenesis of siliciclastic sediments, with implications for provenance studies.

1277 Journal of Geology 104, 525–542.

R
1278 Noblet, C., Lefort, J.P., 1990. Sedimentological evidence for a limited separation between Armorica

SC
1279 and Gondwana during the Early Ordovician. Geology 18, 303–306.

NU
1280 Ó Cofaigh, C., Dowdeswell, J.A., 2001. Laminated sediments in glacimarine environments: diagnostic

1281 criteria for their interpretation. Quaternary Science Reviews 20, 1411–1436.
MA
1282 Oczlon, M.S., Seghedi, A., Carrigan, C.W., 2007. Avalonian and Baltican terranes in the Moesian

1283 Platform (southern Europe, Romania, and Bulgaria) in the context of Caledonian terranes along

1284 the southwestern margin of the East European craton. In: Linnemann, U., Nance, R.D., Kraft,
D

1285
TE

P., Zulauf, G. (Eds.), The Evolution of the Rheic Ocean: From Avalonian–Cadomian Active

1286 Margin to Alleghenian–Variscan Collision. Geological Society of America, Special Paper 423,
P

1287 pp. 375–400.


CE

1288 Oggiano, G., Mameli, P., 2006. Diamictite and oolitic ironstones, a sedimentary association at

1289 Ordovician–Silurian transition in the north Gondwana margin: new evidence from the inner
AC

1290 nappe of Sardinia Variscides (Italy). Gondwana Research 9, 500–511.

1291 Ovenshine, A.T., 1970. Observations of iceberg rafting in Glacier Bay, Alaska, and the identification

1292 of ancient icerafted deposits. Geological Society of America Bulletin 81, 891–894.

1293 Page, A.A., Zalasiewicz, J.A., Williams, M., Popov, L.E., 2007. Were transgressive black shales a

1294 negative feedback modulating glacioeustacy in the Early Palaeozoic icehouse? In: Williams,

1295 M., Haywood, A.M., Gregory, F.J. Schmidt, D.N. (Eds.), Deep-Time Perspectives on Climate

1296 Change: Marrying the Signal from Computer Models and Biological Proxies. The

1297 Micropalaeontological Society, Special Publications, pp. 123–156.


ACCEPTED MANUSCRIPT
52

1298 Paris, F., Bourahrouh, A., Le Hérissé, A., 2000. The effects of the final stages of the Late Ordovician

1299 glaciation on marine palynomorphs (chitinozoans, acritarchs, leiospheres) in well Nl-2 (NE

1300 Algerian Sahara). Review of Palaeobotany and Palynology 113, 87–104.

1301 Passchier, S., Erukanure, E., 2010. Paleoenvironments and weathering regime of the Neoproterozoic

T
Squantum ‘Tillite’, Boston Basin: no evidence of a snowball Earth. Sedimentology 57, 1526–

IP
1302

1303 1544.

R
1304 Passchier, S., Browne, G., Field, B., Fielding, C.R., Krissek, L.A., Panter, K., Pekar, S.F., and

SC
1305 ANDRILL-SMS Science Team, 2011. Early and Middle Miocene Antarctic glacial history

NU
1306 from the sedimentary facies distribution in the AND-2A drill hole, Ross Sea, Antarctica.

1307 Geological Society of America Bulletin 123, 2352–2365.


MA
1308 Perez Loinaze, V.S., Limarino, C.O., Cesari, S.N., 2010. Glacial events in Carboniferous sequences

1309 from Paganzo and Rio Blanco Basins (Northwest Argentina): palynology and depositional

1310 setting. Geologica Acta 8, 399–418.


D

1311
TE

Phillips, E., 2006. Micromorphology of a debris flow deposit: evidence of basal shearing,

1312 hydrofracturing, liquefaction and rotational deformation during emplacement. Quaternary


P

1313 Science Reviews 25, 720–738.


CE

1314 Piçarra, J.M., Robardet, M., Bourahrouh, A., Paris, F., Pereira, Z., Le Menn, J., Gourvennec, R.,

1315 Oliveira, T., Lardeux, H., 2002. Le passage Ordovicien-Silurien et la partie inférieure du
AC

1316 Silurien (Sud-Est du Massif armoricain, France). Comptes Rendus Geoscience 334, 1177–

1317 1183.

1318 Piotrowski, J.A., Larsen, N.K., Menzies, J., Wysota, W., 2006. Formation of subglacial till under

1319 transient bed conditions: deposition, deformation, and basal decoupling under a Weichselian

1320 ice sheet lobe, central Poland. Sedimentology 53, 83–106.

1321 Pohl, A., Donnadieu, Y., Le Hir, G., Ladant, J.-B., Dumas, C., Alvarez-Solas, J., Vandenbroucke,

1322 T.R.A., 2016. Glacial onset predated Late Ordovician climate cooling. Paleoceanography 31,

1323 800–821.

1324 Potter, P.E., Franca, A.B., Spencer, C.W. Caputo, M.V., 1995. Petroleum in glacially-related

1325 sandstones of Gondwana: a review. Journal of Petroleum Geology 18, 397–420.


ACCEPTED MANUSCRIPT
53

1326 Powell, R.D., Cooper, J.M., 2002. A glacial sequence stratigraphic model for temperate, glaciated

1327 continental shelves. In: Dowdeswell, J.A., Ó Cofaigh, C. (Eds.), Glacier-Influenced

1328 Sedimentation on High Latitude Continental Margins: Ancient and Modern. Geological

1329 Society of London, Special Publications 203, pp. 215–244.

T
IP
1330 Rashid, S.A., Ganai, J.A., 2015. Preservation of glacial and interglacial phases in Tethys Himalaya:

1331 evidence from geochemistry and petrography of Permo-Carboniferous sandstones from the

R
1332 Spiti region, Himachal Pradesh, India. Arabian Journal of Geosciences 8, 9345–9363.

SC
1333 Rieu, R., Allen, P.A., Etienne, J.L., Cozzi, A., Wiechert, U., 2006. A Neoproterozoic glacially

NU
1334 influenced basin margin succession and ‘atypical’ cap carbonate associated with bedrock

1335 palaeovalleys, Mirbat area, southern Oman. Basin Research 18, 471–496.
MA
1336 Rieu, R., Allen, P.A., Plotze, M., Pettke, T., 2007. Compositional and mineralogical variations in a

1337 Neoproterozoic glacially influenced succession, Mirbat area, South Oman: implications for

1338 paleoweathering conditions. Precambrian Research 154, 248–265.


D

1339 Robardet, M., 2003. The Armorica ‘microplate’: fact or fiction? Critical review of the concept and
TE

1340 contradictory palaeobiogeographical data. Palaeogeography, Palaeoclimatology, Palaeoecology


P

1341 195, 125–148.


CE

1342 Robardet, M., Doré, F., 1988. The Late Ordovician diamictic formations from southwestern Europe:

1343 North-Gondwana glaciomarine deposits. Palaeogeography, Palaeoclimatology, Palaeoecology


AC

1344 66, 19–31.

1345 Romer, R.L., Hahne, K., 2010. Life of the Rheic Ocean: scrolling through the shale record. Gondwana

1346 Research 17, 236–253.

1347 Rose, C.V., Maloof, A.C., Schoene, B., Ewing, R.C., Linnemann, U., Hofmann, M., Cottle, J.M.,

1348 2013. The end-Cryogenian glaciation of South Australia: Geoscience Canada 40, 256– 293.

1349 Rösel, D., Boger, S.D., Möller, A., Gaitzsch, B., Barth, M., Oalmann, J., Zack, T., 2014. Indo-

1350 Antarctic derived detritus on the northern margin of Gondwana: evidence for continental-scale

1351 sediment transport. Terra Nova 26, 64–71.


ACCEPTED MANUSCRIPT
54

1352 Sabaou, N., Ait-Salem, H., Samy Zazoun, R., 2009. Chemostratigraphy, tectonic setting and

1353 provenance of the Cambro-Ordovician clastic deposits of the subsurface Algerian Sahara.

1354 Journal of African Earth Sciences 55, 158–174.

1355 Sačanski, V., 1993. Boundaries of the Silurian System in Bulgaria on graptolites. Geologica Balcanica

T
IP
1356 23, 25–33.

1357 Sachanski, V., 1994. Age assessment of the Cerecel and Sirman Formations in Sofia Stara planina

R
1358 Mountain (Ordovician, Ashgill). Review of the Bulgarian Geological Society 55, 83–90 (in

SC
1359 Bulgarian with an English abstract)

NU
1360 Sachanski, V., Tenchov, Y., 1993. Lithostratigraphic subdivision of the Silurian sediments in the

1361 Svoge anticline. Review of the Bulgarian Geological Society 54, 71–81 (in Bulgarian with an
MA
1362 English abstract)

1363 Sachanski, V., Tanatsiev, S., 2010. Ordovician–Silurian. In: Angelov, V., Antonov, M. (Eds.).

1364 Explanation Note to the Geological Map of the Republic of Bulgaria scale 1:50,000. Map Sheet
D

1365
TE

Litakovo. Uniscorp, Sofia, pp. 11–15.

1366 Sachanski, V., Göncüoglu, M.C., Gedik, I., 2010. Late Telychian (early Silurian) graptolitic shales and
P

1367 the maximum Silurian highstand in the NW Anatolian Palaeozoic terranes. Palaeogeography,
CE

1368 Palaeoclimatology, Palaeoecology 291, 419–428.

1369 Sachanski, V., Kozlu, H., Göncüoğlu, M.C., 2015. Thuringian affinity of the Silurian–Lower
AC

1370 Devonian succession from the Eastern Taurus, Turkey. Turkish Journal of Earth Sciences 24,

1371 303–324.

1372 Savov, I., Ryan, J., Haydoutov, I., Schijf, J., 2001. Late Precambrian Balkan-Carpathian ophiolite – a

1373 slice of the Pan-African ocean crust? Geochemical and tectonic insights from the Tcherni Vrah

1374 and Deli Jovan massifs, Bulgaria and Serbia. Journal of Volcanology and Geothermal Research

1375 110, 299–318.

1376 Scheffler, K., Buehmann, D., Schwark, L., 2006. Analysis of Late Palaeozoic glacial to postglacial

1377 sedimentary succession in South Africa by geochemical proxies-response to climate evolution

1378 and sedimentary environment. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 184–

1379 203.
ACCEPTED MANUSCRIPT
55

1380 Schönian, F., Egenhoff, S.O., 2007. A Late Ordovician ice sheet in South America: Evidence from the

1381 Cancañiri tillites, southern Bolivia. In: Linnemann, U., Nance, R.D., Kraft, P., Zulauf, G.

1382 (Eds.), The Evolution of the Rheic Ocean: From Avalonian–Cadomian Active Margin to

1383 Alleghenian–Variscan Collision. Geological Society of America, Special Paper 423, pp. 525–

T
IP
1384 548.

1385 Schönlaub, H.P., Ferretti, A., Gaggero, L., Hammarlund, E., Harper, D.A.T., Histon, K., Priewalder,

R
1386 H., Spötl, C., Štorch, P., 2011. The Late Ordovician glacial event in the Carnic Alps (Austria).

SC
1387 In: Gutiérrez-Marco, J.C., Rábano, I., García-Bellido, D. (Eds.), Ordovician of the World.

NU
1388 Madrid. Publicaciones del Instituto Geológico y Minero de España, 14, pp. 515–526.

1389 Scotese, C.R., 2014. Atlas of Silurian and Middle-Late Ordovician Paleogeographic Maps (Mollweide
MA
1390 Projection), Maps 73–80, Volume 5, The Early Paleozoic, PALEOMAP Atlas for ArcGIS,

1391 PALEOMAP Project, Evanston, IL.

1392 Shaw, J., Gutierrez-Alonso, G., Johnston, S.T., Galan, D.P., Pastor Galan, D., 2014. Provenance
D

1393
TE

variability along the Early Ordovician north Gondwana margin: Paleogeographic and tectonic

1394 implications of U-Pb detrital zircon ages from the Armorican Quartzite of the Iberian Variscan
P

1395 belt. Geological Society of America Bulletin 126, 702–719.


CE

1396 Spassov, H., 1960. Stratigraphie des Ordoviziums und Silurs im Kern der Svoge-Antiklinale. Travaux

1397 sur la Geologie de Bulgarie, Serie Stratigraphie et Tectonique 1, 161–202 (in Bulgarian with
AC

1398 Russian and German abstracts)

1399 Spassov, H., 1968. Ordovician and Silurian. In: Tzankov, V., Spassov, H. (Eds.), Stratigraphy of

1400 Bulgaria. Nauka i Izkustvo, Sofia, pp. 75–100.

1401 Squire, R.J., Campbell, I.H., Allen, C.M., Wilson, C.J.L., 2006. Did the Transgondwanan

1402 Supermountain trigger the explosive radiation of animals on Earth? Earth and Planetary

1403 Science Letters 250, 116–133.

1404 Stampfli, G.M., Hochard, C., Vérard, C., Wilhem, C., von Raumer, J., 2013. The formation of Pangea.

1405 Tectonophysics 593, 1–19.


ACCEPTED MANUSCRIPT
56

1406 Steiner, J., Falk, F., 1981. The Ordovician Lederschiefer of Thuringia. In: Hambrey, M.J., Harland,

1407 W.B. (Eds.), Earth’s Pre-Pleistocene Glacial Record. Cambridge University Press, Cambridge,

1408 pp. 579–581.

1409 Stickley, C.E., St. John, K., Koç, N., Jordan, R.W., Passchier, S., Pearce, R.B., Kearns, L.E., 2009.

T
IP
1410 Evidence for middle Eocene Arctic sea ice from diatoms and ice-rafted debris. Nature 460,

1411 376–379.

R
1412 St. John, K., Passchier, S., Tantillo, B., Darby, D., Kearns, L., 2015. Microfeatures of modern sea ice-

SC
1413 rafted sediment and implications for paleo-sea-ice reconstructions. Annals of Glaciology 56,

NU
1414 83–93.

1415 Štorch, P., 1990. Upper Ordovician – Lower Silurian sequences of the Bohemian Massif, central
MA
1416 Europe. Geological Magazine 127, 225–239.

1417 Štorch, P., 2006. Facies development, depositional settings and sequence stratigraphy across the

1418 Ordovician-Silurian boundary: a new perspective from the Barrandian area of the Czech
D

1419
TE

Republic. Geological Journal 41, 163–192.

1420 Štorch, P., Schönlaub, H.P., 2012. Ordovician-Silurian boundary graptolites of the Southern Alps,
P

1421 Austria. Bulletin of Geosciences 87, 755–766.


CE

1422 Suttner, L.J., Basu, A., Mack, G.H., 1981. Climate and the origin of quartz arenites. Journal of

1423 Sedimentary Petrology 51, 235–246.


AC

1424 Tachibana, T., 2013. Lonestones as indicators of tsunami deposits in deep-sea sedimentary rocks of

1425 the Miocene Morozaki Group, central Japan. Sedimentary Geology 289, 62–73.

1426 Tait, J.A., Bachtadse, V., Franke, W., Soffel, H.C., 1997. Geodynamic evolution of the European

1427 Variscan fold belt: Palaeomagnetic and geological constraints. Geologische Rundschau 86,

1428 585–598.

1429 Talling, P.J., Masson, D.G., Sumner, E.J., Malgesini, G., 2012. Subaqueous sediment density flows:

1430 depositional processes and deposit types. Sedimentology 59, 1937–2003.

1431 Taylor, J., Dowdeswell, J.A., Kenyon, N.H., Ó Cofaigh, C., 2002. Late Quaternary architecture of

1432 trough-mouth fans: debris flows and suspended sediments on the Norwegian margin. In:
ACCEPTED MANUSCRIPT
57

1433 Dowdeswell, J.A., Ó Cofaigh, C. (Eds.), Glacier-Influenced Sedimentation on High Latitude

1434 Continental Margins. Geological Society of London, Special Publications 203, pp. 55–71.

1435 Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: Its Composition and Evolution.

1436 Blackwell, Oxford.

T
IP
1437 Thomas, G.S.P., Connell, R.J., 1985. Iceberg drop, dump and grounding structures from Pleistocene

1438 glacio-lacustrine sediments, Scotland. Journal of Sedimentary Petrology 55, 243–249.

R
1439 Tomkins, J.D., Lamoureux, S.F., Antoniades, D., Vincent, W.F., 2009. Sedimentary pellets as an ice-

SC
1440 cover proxy in a High Arctic ice-covered lake. Journal of Paleolimnology 41, 225–242.

NU
1441 Torsvik, T.H., Cocks, L.R.M., 2011. The Palaeozoic palaeogeography of central Gondwana. In: Van

1442 Hinsbergen, D.J.J., Buiter, S.J.H., Torsvik, T.H., Gaina, C., Webb, S.J. (Eds.), The Formation
MA
1443 and Evolution of Africa: A Synopsis of 3.8 Ga of Earth History. Geological Society of London,

1444 Special Publications 357, pp. 137–166.

1445 Troedson, A.L., Riding, J.B., 2002. Upper Oligocene to lowermost Miocene strata of King George
D

1446
TE

Island, South Shetland Islands, Antarctica: stratigraphy, facies analysis, and implications for

1447 the glacial history of the Antarctic Peninsula. Journal of Sedimentary Research 72, 510–523.
P

1448 Turner, B.R., Makhlouf, I.M., Armstrong, H.A., 2005. Late Ordovician (Ashgillian) glacial deposits in
CE

1449 southern Jordan. Sedimentary Geology 181, 73–91.

1450 van de Kamp, P., 2010. Arkose, subarkose, quartz sand, and associated muds derived from felsic
AC

1451 plutonic rocks in glacial to tropical humid climates. Journal of Sedimentary Research 80, 895–

1452 918.

1453 van der Meer, J.J.M., 1996. Micromorphology. In: Menzies, J. (Ed.), Past Glacial Environments –

1454 Sediments, Forms and Techniques. Butterworth-Heinemann, Oxford, pp. 335–355.

1455 Vesely, F.F., Assine, M.L., 2006. Deglaciation sequences in the Permo-Carboniferous Itararé Group,

1456 Paraná Basin, southern Brazil. Journal of South American Earth Sciences 22, 156–168.

1457 Visser, J.N.J., 1994. The interpretation of massive rain-out and debris-flow diamictites from the

1458 glacial marine environment. In: Deynoux, M., Miller, J.M.G., Domack, E.W., Eyles, N.,

1459 Fairchild, I.J., Young, G.M. (Eds.), Earth's Glacial Record. Cambridge University Press,

1460 Cambridge, pp. 83–94.


ACCEPTED MANUSCRIPT
58

1461 Visser, J.N.J., 1997. Deglaciation sequences in the Permo-Carboniferous Karoo and Kalahari basins of

1462 Southern Africa: a tool in the analysis of cyclic glaciomarine basin fills. Sedimentology 44,

1463 507–521.

1464 Visser, J.N.J., Young, G.M., 1990. Major element geochemistry and paleoclimatology of the Permo–

T
IP
1465 Carboniferous glacigene Dwyka Formation and post–glacial mudrocks in Southern Africa.

1466 Palaeogeography, Palaeoclimatology, Palaeoecology 81, 49–57.

R
1467 von Eynatten, H., Tolosana-Delgado, R., Karius, V., 2012. Sediment generation in modern glacial

SC
1468 settings: grain-size and source-rock control on sediment composition. Sedimentary Geology

NU
1469 280, 80–92.

1470 von Quadt, A., Peycheva, I., Haydutov, I., 1998. U–Zr dating of Tcherny Vrach metagabbro, West
MA
1471 Balkan, Bulgaria. Compte Rendus de l'Academie Bulgare des Sciences 51, 81–84.

1472 Winchester, J.A., Pharaoh, T.C., Verniers, J., Ioane, D., Seghedi, A., 2006. Palaeozoic accretion of

1473 Gondwana-derived terranes to the East European Craton: Recognition of detached terrane
D

1474
TE

fragments dispersed after collision with promontories. In: Gee, D.G., Stephenson, R. (Eds.),

1475 European Lithosphere Dynamics. Geological Society of London, Memoirs 32, pp. 323–332.
P

1476 Yanev, S., 1993. Gondwana Palaeozoic terranes in the Alpine collage system on the Balkans. Journal
CE

1477 of Himalayan Geology 4, 257–270.

1478 Yanev, S., 1997. Palaeozoic migration of terranes from the basement of the eastern part of the Balkan
AC

1479 peninsula from peri-Gondwana to Laurussia. In: Göncüoğlu, M.C., Derman, A.S. (Eds.), Early

1480 Palaeozoic Evolution in NW Gondwana. Turkish Association of Petroleum Geologists, Special

1481 Publication 3, pp. 89–100.

1482 Yanev, S., 2000. Palaeozoic terranes of the Balkan Peninsula in the framework of Pangea assembly.

1483 Palaeogeography, Palaeoclimatology, Palaeoecology 161, 151–177.

1484 Yanev, S., Tzankov, Tz., Boncheva, I., 1995. Lithostratigraphy and Late Alpine structure of the

1485 Palaeozoic terrains in the Shipka Part of Stara Planina Mountains. Geologica Balcanica 25, 3–

1486 26.
ACCEPTED MANUSCRIPT
59

1487 Yanev, S., Lakova, I., Boncheva, I., Sachanski, V., 2005. The Moesian and Balkan Terranes in

1488 Bulgaria: Palaeozoic marine basin development, palaeogeography and tectonic evolution.

1489 Geologica Belgica 8, 185–192.

1490 Yanev, S., Goncuoğlu, M.C., Gedik, I., Lakova, I., Boncheva, I., Sachanski, V., Okuyucu, C., Ozgul,

T
IP
1491 N., Timur, E., Malyakov, Y., Saydam, G., 2006. Stratigraphy, correlations and

1492 palaeogeography of Palaeozoic terranes of Bulgaria and NW Turkey: a review of recent data.

R
1493 In: Robertson, A.H.F., Mountrakis, D. (Eds.), Tectonic Development of the Eastern

SC
1494 Mediterranean Region. Geological Society of London, Special Publications 260, pp. 51–67.

NU
1495 Young, G.M., 2001. Comparative geochemistry of Pleistocene and Paleoproterozoic (Huronian)

1496 glaciogenic laminated deposits: relevance to crustal and atmospheric composition in the last 2.3
MA
1497 Ga. Journal of Geology 109, 463–477.

1498 Young, G.M., Nesbitt, H.W., 1999. Paleoclimatology and provenance of the glaciogenic Gowganda

1499 Formation (Paleoproterozoic), Ontario, Canada: a chemostratigraphic approach. Geological


D

1500
TE

Society of America Bulletin 111, 264–274.

1501 Young, G.M., Minter, W.E.L., Theron, J.N., 2004. Geochemistry and palaeogeography of upper
P

1502 Ordovician glaciogenic rocks in the Table Mountain Group, South Africa. Palaeogeography,
CE

1503 Palaeoclimatology, Palaeoecology 214, 323–345.

1504 Ziegler, P.A., Wimmenauer, W., 2001. Possible glaciomarine diamictites in Lower Paleozoic series of
AC

1505 the southern Black Forest (Germany). Neues Jahrbuch für Geologie und Paläontologie,

1506 Monatshefte 2001, 500–512.

1507
ACCEPTED MANUSCRIPT
60

T
IP
R
SC
NU
MA

1508
D

1509 Figure 1
TE

1510
P
CE
AC
ACCEPTED MANUSCRIPT
61

T
IP
R
SC
NU
MA
1511
1512 Figure 2
D

1513
PTE
CE
AC
ACCEPTED MANUSCRIPT
62

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1514
1515 Figure 3

1516
ACCEPTED MANUSCRIPT
63

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1517
1518 Figure 4

1519
ACCEPTED MANUSCRIPT
64

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1520
1521 Figure 5

1522
ACCEPTED MANUSCRIPT
65

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1523
1524 Figure 6

1525
ACCEPTED MANUSCRIPT
66

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1526
1527 Figure 7

1528
ACCEPTED MANUSCRIPT
67

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1529
1530 Figure 8

1531
ACCEPTED MANUSCRIPT
68

T
IP
R
SC
NU
MA
D
PTE
CE
AC

1532
1533 Figure 9

1534
ACCEPTED MANUSCRIPT
69

T
IP
R
SC
NU
MA
D
TE

1535
P

1536 Figure 10
CE

1537
AC
ACCEPTED MANUSCRIPT
70

T
IP
R
SC
NU
MA
D

1538
TE

1539 Figure 11

1540
P
CE
AC
ACCEPTED MANUSCRIPT
71

T
IP
R
SC
NU
MA
D
TE

1541
P

1542 Figure 12
CE

1543
AC
ACCEPTED MANUSCRIPT
72

T
IP
R
SC
NU
MA
D
TE

1544
P

1545 Graphical abstract


CE

1546
AC
ACCEPTED MANUSCRIPT
73

1547 FIGURE CAPTION

1548

1549 Fig. 1. Palaeozoic terranes and exposures (shown as blue patches) in Bulgaria (from

T
1550 Yanev, 2000; Yanev et al., 2006).

IP
1551

R
1552 Fig. 2. Outcrops of Hirnantian glaciomarine rocks (Sirman Fm) in the western

SC
1553 Srednogorie Zone (data used from Angelov et al., 2010, 2011). Symbols: 1 – Quaternary

1554 sediments, 2 – Permian volcanic and sedimentary rocks, 3 – Carboniferous volcanic and

1555
NU
sedimentary rocks, 4 – Lower–Upper Devonian sedimentary rocks, 5 – Pridoli–Lower

Devonian sedimentary rocks, 6 – Llandovery–Pridoli sedimentary rocks, 7 – Saltar Fm


MA
1556

1557 (Upper Ordovician–Llandovery) and Sirman Fm (Upper Ordovician), 8 – Tseretsel Fm,

1558 Grohoten Fm and siltstone-argillite metaformation (Middle–Upper Ordovician), 9 – studied


D
TE

1559 sections. Inset map: Alpine tectonic subdivision of Bulgaria (from Ivanov, 1988) with

1560 designated location of the study area. Hachures outline the Srednogorie Zone and dotted
P

1561 sector corresponds to the Svoge Unit.


CE

1562
AC

1563 Fig. 3. (left) Stratigraphy of the Ordovician System in the western Srednogorie Zone

1564 (from Sachanski and Tanatsiev, 2010) with only formal lithostratigraphic units shown. (right)

1565 Lithologic log of the Sirman Fm and vertical distribution of the recognized glaciomarine

1566 lithofacies.

1567

1568 Fig. 4. (a) Sharp planar contact (marked by hammer) between fissile mudstones of the

1569 Tseretsel Fm (TsF) and thick-bedded diamictites of the Sirman Fm (SmF). Scale: hammer is

1570 31 cm long; (b) Slightly undulating contact (arrows) between the same two units. (c) Sharp

1571 conformable contact (marked by hammerhead) between laminated mudstones of the Sirman
ACCEPTED MANUSCRIPT
74

1572 Fm (SmF) and thin-bedded cherts of the Saltar Fm (StF); (d) Thick beds of structureless

1573 diamictites separated by thin mudstone bed (marked by hammerhead). Scale: hammer is 33

1574 cm long; (e) Slump developed in diamictite; (f) Lower and upper parts of the Sirman Fm

T
1575 (boundary marked by dashed line) consisting of thick-bedded massive diamictites (md) and

IP
1576 medium-bedded laminated diamictites (ld), respectively. Scale: yardstick (arrow) is 1 m long.

R
1577

SC
1578 Fig. 5. Structureless diamictites (lithofacies 1) from the lower part (a–d) and laminated

1579 rocks from the upper part of the Sirman Fm (e, f): (a) Well rounded pebbly lonestones

1580
NU
(encircled) probably derived from mudstones of the underlying Tseretsel Fm. Coin diameter
MA
1581 is 24 mm; (b) Elongated outsized clasts of mudstones (smaller one shown by arrow) with

1582 subvertical orientation to bedding (dashed line indicates bedding plane); (c) Pebble-sized

1583 lonestone of sandstone lying with its long axis subparallel to bedding. Pen is 14 cm long; (d)
D
TE

1584 Rock specimen with flat-shaped pebbly intraclasts of dark grey mudstones (arrows) having

1585 various sizes and degree of roundness; (e) Polished slab of crudely laminated diamictite
P

1586 (lithofacies 3) showing mudstone extraclasts (blue arrows) and intraclasts (yellow arrows)
CE

1587 with maximum granule size, mostly elongated shape and subparallel orientation to bedding.
AC

1588 Some authigenic pyrite crystals with developed whitish quartz strain fringes are encircled.

1589 Coin diameter is 11 mm; (f) Polished slab of finely laminated mudstone (lithofacies 4).

1590

1591 Fig. 6. Various types of clast-poor and clast-rich diamictites distinguished according to

1592 the sand/mud ratio (i.e., based on the textural classification of poorly sorted sediments from

1593 Moncrieff, 1989; modified by Hambrey, 1994): (a, b) Muddy diamictites (mi – mudstone

1594 intraclast, arrow – muscovite); (c, d) Intermediate diamictites; (e, f) Sandy diamictites (circle

1595 – plagioclase, arrows – pressure-solution grain contacts, dashed line – grain lineation). Note

1596 the poor sorting and various degree of roundness of the sand-sized grains which are almost
ACCEPTED MANUSCRIPT
75

1597 entirely represented by detrital quartz. All microphotographs are in cross-polarized light.

1598 Scale bar is 0.3 mm across.

1599

T
1600 Fig. 7. Detrital grains of mono- and polycrystalline quartz (a–f), quartz-rich rock

IP
1601 fragments (g–k), and feldspar (l) in the diamictites: (a) Grain with undulatory extinction

R
1602 (arrow); (b) Three grains showing deformation lamellae; (c) Volcanic quartz with partly

SC
1603 preserved idiomorphic crystal habit, rounded edges and embayments; (d) Sedimentary quartz

1604 with abraded quartz overgrowths (arrows); (e, f) Polycrystalline grains with mostly elongated

1605
NU
individual crystals and sutured crystal boundaries (arrow – detrital zircon); (g) Quartzite
MA
1606 fragment showing evidence of pressure solution (concavo-convex to slightly serrated grain

1607 boundaries) in the parent rock; (h) Extrabasinal clast of very fine grained quartzose sandstone;

1608 (i) Lithic grain consisting of microcrystalline quartz (probably derived from chert); (j) Exotic
D
TE

1609 rock fragment of granitoid (plagioclase shown by arrow); (k) Extraclast of fine grained

1610 arkosic sandstone; (l) Plagioclase grains (arrows) with the upper one showing replacement by
P

1611 clay minerals. All microphotographs except c (in plane-polarized light) are in cross-polarized
CE

1612 light. Scale bar is 0.3 mm across.


AC

1613

1614 Fig. 8. (a) Brownish mudstone extraclasts with thin oxidized dark rims greatly

1615 resembling rocks of the underlying Tseretsel Fm; (b) Associated extrabasinal (ex) and

1616 intrabasinal (in) clasts of mudstones showing different sizes, shapes, colours, and degree of

1617 roundness. The extraclast has oxidized rim and the intraclast is affected by mechanical

1618 compaction; (c) Extrabasinal mudstone clast with quartz veinlet inherited from the parent

1619 rock; (d–f) Intraclasts (arrows) derived from semi-consolidated mudstones (see also Figure

1620 9f) displaying good rounding, local effects of mechanical compaction, and absence of sharply

1621 outlined boundaries or oxidized rims; (g) Elongated rip-up clast of diamictite with distinct
ACCEPTED MANUSCRIPT
76

1622 boundaries; (h–j) Sediment aggregates (shown by dashed lines) having irregular shapes,

1623 diffuse boundaries, and higher amount of mud compared to the surrounding diamictite

1624 groundmass. These are interpreted as ‘till pellets’, i.e., frozen masses of ice-rafted debris. All

T
1625 microphotographs are in plane-polarized light. Scale bar is 0.3 mm across.

IP
1626

R
1627 Fig. 9. Depositional and deformational fabrics in the diamictites (a–e, i–k) and

SC
1628 mudstones (f-h): (a, b) Galaxy/turbate structure (dashed circles) outlined by circular

1629 alignment of quartz grains and matrix (lithofacies 1); (c) Rotational structure (arrows) around

1630
NU
large core grain (lithofacies 1); (d) Soft deformation of the matrix (arrows) below elongated
MA
1631 quartz grain with preserved steep dip of its long axis (lithofacies 3); (e) Syndepositionally

1632 disturbed matrix (arrows) below granule-sized detrital grain (lithofacies 3); (f, g)

1633 Homogeneous microfabrics consisting of low birefringent clay particles and randomly
D
TE

1634 dispersed silt grains dominated by quartz (lithofacies 2); (h) Finely laminated fabric

1635 (lithofacies 4) with quartz silt enrichment in the light coloured laminae. The latter are planar
P

1636 to slightly undulating and only locally discontinuous without evidence of significant soft
CE

1637 deformation or erosion of the underlying laminae; (i) Parallel lamination defined by vertical
AC

1638 changes in the colour, grain size and sand/mud ratio (lithofacies 3); (j) Chamosite strain

1639 fringes (arrows) developed between detrital quartz grains (lithofacies 3); (k) Quartz strain

1640 fringes (arrows) grown around diagenetic pyrite crystal (lithofacies 3). All microphotographs

1641 except a, g and k (in cross-polarized light) are in plane-polarized light. Scale bar is 0.3 mm

1642 across.

1643

1644 Fig. 10. Schematic representation of the successive deglaciation phases characterized by

1645 deposition of three glaciomarine lithofacies in low-energy mid- to outer-shelf setting:

1646 dominantly structureless diamictites with lonestones (A), crudely laminated diamictites (B),
ACCEPTED MANUSCRIPT
77

1647 and finely laminated mudstones (C). Progressive retreat of the ice front (grounded or floating

1648 ice sheet) and related sea-level rise resulted in waning of mass-flow processes, decreasing

1649 supply of ice-rafted debris (IRD), increasing contribution from suspension settling, and

T
1650 reduced sedimentation rates.

IP
1651

R
1652 Fig. 11. A-CN-K diagram with plotted molar proportions of Al2O3, CaO*+Na2O and

SC
1653 K2O for the analyzed samples of diamictites (red circles) and mudstones (yellow circles).

1654 Data points outline a cluster that closely follows the ideal weathering trend (arrows) of felsic

1655
NU
rocks (Nesbitt and Young, 1984; Fedo et al., 1995), thus implying that the glaciomarine
MA
1656 sediments were derived from chemically weathered parent rocks having granodioritic to

1657 granitic composition. However, the compositional and some textural maturity of the

1658 diamictites indicates recycling of older sediments or sedimentary rocks (see text).
D
TE

1659 Abbreviations: CIA – chemical index of alteration, pl – plagioclase, kf – potassium feldspar,

1660 sm – smectite, il– illite, ms – muscovite, ka – kaolinite, gi – gibbsite, ch – chlorite.


P

1661
CE

1662 Fig. 12. (A) Present-day geographical distribution of Hirnantian glaciomarine rocks
AC

1663 deposited along the periphery of the North Gondwana platform. The occurrences in Morocco,

1664 (Le Heron et al., 2007, 2008), Algeria (Beuf et al., 1971; Paris et al., 2000), Portugal

1665 (Brenchley et al., 1991; Couto et al., 2013), Spain (Fortuin, 1984; Robardet and Doré, 1988;

1666 Bernárdez et al., 2006; Álvaro and Van Vliet-Lanoë, 2009; Gutiérez-Marco et al., 2010),

1667 Turkey (Monod et al., 2003; Ghienne et al., 2010) and Iran (Ghavidel-syooki et al., 2011)

1668 indicate sedimentation close to the maximum ice-front position of the Gondwana-based ice

1669 sheet (or satellite ice caps). The rest occurrences in France, Germany, Italy, Austria, Czech

1670 Republic, Serbia (see references in text) and Bulgaria comprise ice-distal deposits. (B)

1671 Palaeogeographic reconstruction of North Gondwana for the Hirnantian age (after Scotese,
ACCEPTED MANUSCRIPT
78

1672 2014) with inferred location of the Balkan Terrane and juxtaposition of the other peri-

1673 Gondwana terranes where dominantly glaciomarine sedimentation occurred (compiled from

1674 Monod et al., 2003; Yanev et al., 2006; Linnemann et al., 2011; Torsvik and Cocks, 2011;

T
1675 Melchin et al., 2013; Stampfli et al., 2013; Sachanski et al., 2015, and references therein).

IP
1676

R
SC
NU
MA
D
P TE
CE
AC
ACCEPTED MANUSCRIPT
79

1677 Table 1

1678 Major oxides (weight %) and calculated CIA values of the Hirnantian glaciomarine

1679 rocks.

T
1680 d i a m i c t i t e s m u d s t o n e s*

IP
1681 _____________________________________________ __________________________

1682 __________________________________________________________________________________

R
1683 ___________

SC
1684 SiO2 69.13 73.93 81.84 72.32 80.94 76.91 76.02 81.23 54.59 53.81

1685 62.59 34.44

1686
1687
TiO2

0.64 0.67
0.70 0.47 0.53 0.52 0.57
NU
0.60 0.63 0.63 1.38 0.99
MA
1688 Al2O3 9.36 10.37 8.94 13.68 9.37 10.72 8.62 8.10 20.88 20.61

1689 13.22 8.89


D

1690 Fe2O3 8.40 4.26 2.21 3.62 1.99 3.33 6.29 3.50 8.66 9.35
TE

1691 6.83 23.26

1692 MnO 0.37 0.18 0.01 0.01 0.01 0.01 0.02 0.03 0.04 0.05
P

1693 0.26 3.65


CE

1694 MgO 2.67 1.74 1.10 1.37 0.73 0.99 1.18 1.10 2.68 2.81

1695 4.38 6.19


AC

1696 CaO 1.70 1.72 0.59 0.60 0.67 0.64 0.73 0.78 0.89 0.65

1697 2.07 3.02

1698 Na2O 0.68 1.35 0.83 1.47 0.12 0.84 1.00 0.93 1.82 0.65

1699 1.09 0.53

1700 K2O 1.33 1.54 1.78 2.69 2.35 2.40 2.32 1.35 4.30 4.41

1701 2.66 0.83

1702 P2O5 0.87 0.24 0.22 0.21 0.32 0.27 0.19 0.17 0.18 0.27

1703 0.23 1.24

1704 L.O.I. 4.88 3.99 2.02 3.34 2.56 2.98 3.10 2.01 4.55 6.05

1705 5.73 16.77


ACCEPTED MANUSCRIPT
80

1706 Total 100.09 99.79 100.07 99.83 99.63 99.69 100.10 99.83 99.97 99.65

1707 99.70 99.49

1708 CIA 72 63 70 69 76 70 63 67 70 76

1709 67 76

T
1710 __________________________________________________________________________________

IP
1711 _____________

R
1712 *The last sample is from finely laminated mudstones forming the top of Sirman Fm and the other

SC
1713 three mudstone samples are from non-laminated rocks in the lower part of the unit.

1714

NU
MA
D
P TE
CE
AC
ACCEPTED MANUSCRIPT
81

1715 Highlights
1716  Deposition occurred on the non-glaciated shelf of the North Gondwana platform
1717  Ice-intermediate and ice-distal deposits constitute a typical deglaciation sequence
1718  Provenance was related to sedimentary recycling of mature Cambro-Ordovician sands

T
R IP
SC
NU
MA
D
P TE
CE
AC

You might also like