<<

1 Graptolites as fossil geo-thermometers and source material of

2 hydrocarbons: an overview of four decades of progress

3 Qingyong Luoa, b, Goodarzi Fariborzc, Ningning Zhonga, b*, Ye Wanga, b, Nansheng Qiua, b,

4 Christian B. Skovstedd, Václav Suchýe, Niels Hemmingsen Schovsbof, Rafał Morgag,

5 Jingyue Haoa, b, Anji Liua, b, Jin Wua, b, Xu Mina, b, Weixun Caoa, b, Jia Wua, b

6 a State Key Laboratory of Petroleum Resources and Prospecting, China University of Petroleum, Beijing 102249, China;

7 b College of Geoscience, China University of Petroleum, Beijing 102249, China;

8 c FG &Partner Ltd, Research Group, 29 Hawkside Mews NW., Calgary, Alberta, ;

9 d Department of Palaeobiology, Swedish Museum of Natural History, Box 50007, SE-104 05 Stockholm, Sweden;

10 e Nuclear Physics Institute, v. v. i.,Academy of Sciences of the Czech Republic, Na Truhlářce 39/64, 180 86 Prague 8, Czech

11 Republic;

12 f Geological Survey of Denmark and Greenland (GEUS), Øster Voldgade 10, DK-1350 Copenhagen K, Denmark;

13 g Silesian University of Technology, Faculty of Mining, Safety Engineering and Industrial Automation, Institute of Applied

14 Geology, Akademicka 2, 44-100 Gliwice, Poland.

15

16 *Corresponding author at: State Key Laboratory of Petroleum Resources and Prospecting, China

17 University of Petroleum, Changping, Beijing, 102249, China. Tel.: +86 10 89734548. E-mail address:

18 [email protected].

19

20 Abstract

21 The thermal maturity of Lower Paleozoic graptolite-bearing marine sediments, which host many

22 hydrocarbon deposits worldwide, has long been difficult to determine due to the absence of wood-derived

23 vitrinite particles for conventional vitrinite reflectance. In 1976, graptolite reflectance was introduced as

24 a new indicator for organic maturity of these deposits and has been used since in many regional studies.

25 The majority of these studies, however, were done on a limited sample set and a limited range of thermal

26 maturity, which resulted in a number of controversial views concerning the usefulness of graptolite

27 reflectance as an alternative paleothermal indicator and its correlation with vitrinite reflectance through

28 various proxies. In this paper, we review previous studies and combine those analyses with new data to 29 assess the physical and chemical characteristics of graptolite periderm with increasing thermal maturity.

30 We conclude that graptolite random reflectance (GRor) is a better parameter for the thermal maturity

31 assessment than graptolite maximum reflectance (GRomax) due to the better quality of available data and

32 time saving. Combining published data with results of our study of both natural and heat-treated

33 graptolites and vitrinite, we present a new correlation between GRor and equivalent vitrinite reflectance

34 (EqVRo), as EqVRo = 0.99GRor + 0.08. Chemical composition of graptolite periderm is similar to vitrinite;

35 graptolites are mainly kerogen Type II-III, are gas prone and have a substantial hydrocarbon potential.

36 Lower Paleozoic graptolite-bearing organic-rich sediments are important gas source rocks and

37 reservoirs globally and make a significant contribution to worldwide petroleum reserves.

38 Keywords: Optical characteristics; Graptolite reflectance; Chemical composition; Microstructure;

39 Wufeng–Longmaxi Formations; Alum Shale; Hot shale; .

40

41 1. Introduction and previous studies

42 Lower Paleozoic graptolite-bearing rocks were mainly deposited in marine environments, and are

43 important source rocks globally, especially shale gas deposits in the Wufeng–Longmaxi (also known as

44 Wufeng–Lungmachi) sediments from China (Zou et al., 2010; Zou et al., 2012; Dai et al., 2014; Dai et

45 al., 2016; Luo et al., 2016; Zou et al., 2016; Luo et al., 2017; Luo et al., 2018). These organic matter

46 (OM)-rich facies were also identified as true or potential hydrocarbon source rocks in the Anadarko basin

47 in the USA (Wang and Philp, 1997), North Africa and Arabian Peninsula (Jones and Stump, 1999; Lüning

48 et al., 2000), Taurus region of Turkey (Varol et al., 2006), the Czech Republic (Suchý et al., 2002), and

49 the Siberian platform (Makarov and Bazhenova, 1981). More recently, such have been recognized

50 as potential targets for unconventional shale gas deposits in the Norwegian-Danish Basin (Schovsbo et al.

51 2011, 2014; Pool et al. 2012), and Baltic basin of Central Europe (Littke et al., 2011; Karcz et al., 2013;

52 Yang et al. 2017).

53 Vitrinite reflectance is a most commonly used indicator for the thermal maturity (Stach et al., 1982;

54 Taylor et al., 1998; Suárez-Ruiz et al., 2012; Hackley and Cardott, 2016). However, due to the lack of

55 vitrinite (coalified wood) in pre- rocks, the determination of thermal maturity of Lower

56 Paleozoic graptolite-bearing rocks is always a difficult topic and a hot debate for petroleum industry

57 (Goodarzi, 1984; Goodarzi, 1985a; Goodarzi and Norford, 1985, 1987; Bertrand and Heroux, 1987;

58 Bustin et al., 1989; Bertrand, 1990; Goodarzi et al., 1992a; Bertrand, 1993; Petersen et al., 2013; Luo et 59 al., 2016; Luo et al., 2017; Luo et al., 2018). Thus, the surrogate proxies, such as the reflectance of

60 zooclasts, vitrinite-like particles and solid bitumen, Tmax, and biomarkers, have been proposed to assess

61 organic maturity in Lower Paleozoic graptolite-bearing sediments (Teichmüller, 1978; Goodarzi and

62 Norford, 1987; Jacob, 1989; Schoenherr et al., 2007; Suárez-Ruiz et al., 2012; Schmidt et al., 2019).

63 Tmax may be unreliable due to low S2 in overmature sediments (Peters, 1986; Peters and Cassa, 1994).

64 The maturity-related biomarker ratios may be also influenced by depositional environments and biological

65 sources, which will increase the difficulty of data interpretation (Radke and Welte, 1983; Radke et al.,

66 1986; Radke, 1988; George and Ahmed, 2002; Peters et al., 2005). In addition, the biomarker ratios may

67 be invalid to assess thermal maturity of overmature sediments (Peters et al., 2005). Graptolite-bearing

68 rocks often contain bitumen and other zooclasts (Teichmüller, 1978; Goodarzi and Norford, 1987; Jacob,

69 1989; Schoenherr et al., 2007; Suárez-Ruiz et al., 2012; Schmidt et al., 2019). The difficulty with using

70 of solid bitumen reflectance is due to large variation in most samples, and their origin, e.g., by thermal

71 cracking, biodegradation and deasphalting (George et al., 1994; Hwang et al., 1998; Mastalerz et al., 2018),

72 all of which increase difficulty in data interpretation (e.g., Gonçalves et al., 2014; Fink et al., 2016). The

73 origin of discrete vitrinite-like particles remains controversial, and possible explanations includes:

74 migrated bitumen, either indigenous or exogenous to the host rock (Bertrand and Heroux, 1987);

75 gelification of polysaccharides (Buchardt and Lewan, 1990); residues of algae after maturation (Wang et

76 al., 1994); biodegraded zooclasts that are the product of a reducing to strongly reducing environment

77 (Xiao et al., 1997); marine humification of planktonic and benthic organisms (Romankevich, 1984), the

78 so called “marine vitrinite group” (Zhong and Qin, 1995); and fragments of graptolites (Petersen et al.,

79 2013).

80 Zooclasts have clear advantage in reflectance studies due to their specific biological sources

81 compared to that of solid bitumen and vitrinite-like particles, and thus their reflectance was naturally

82 regarded as having a superior potential as a thermal maturity proxy. In general, graptolites are more

83 common than other zooclasts (e.g., chitinozoans and scolecodonts) in Lower Paleozoic marine rocks, and

84 as a result, the nature of graptolite reflectance has been a “hot topic” for organic petrologists over several

85 decades (Kurylowicz et al., 1976; Teichmüller, 1978; Goodarzi, 1984, 1985a; Bertrand and Heroux, 1987;

86 Bertrand, 1990; Link et al., 1990; Cardott and Kidwai, 1991; Hoffknecht, 1991; Goodarzi et al., 1992b;

87 Malinconico, 1992; Tricker et al., 1992; Malinconico, 1993; Wang et al., 1993; Cole, 1994; Gentzis et al.,

88 1996; Liu et al., 2001; Bertrand et al., 2003; Petersen et al., 2013; İnan et al., 2016; Lavoie et al., 2016; 89 Luo et al., 2016; Luo et al., 2017; Luo et al., 2018; Synnott et al., 2018; Wang et al., 2019a).

90 Graptolite reflectance is a useful proxy of thermal maturity (Goodarzi, 1984; Goodarzi and Norford,

91 1985), and has been used to assess eroded thicknesses when used in conjunction with reflectance of

92 bitumen and sedimentological and tectonic evidences (Goodarzi et al., 1992b; Gentzis et al., 1996). The

93 graptolite maximum reflectance (GRomax) or graptolite random reflectance (GRor) was adopted to estimate

94 thermal maturity in worldwide sediments (Goodarzi, 1984; Goodarzi and Norford, 1985; Goodarzi et al.,

95 1985; Goodarzi and Norford, 1989; Goodarzi et al., 1992b; Malinconico, 1992, 1993; Rantitsch, 1995;

96 Gentzis et al., 1996; Bertrand et al., 2003; Petersen et al., 2013; İnan et al., 2016; Lavoie et al., 2016; Luo

97 et al., 2016; Luo et al., 2017; Luo et al., 2018). Some researchers have established correlations among

98 reflectances of graptolite and other zooclasts, pyrobitumen and vitrinite (Bertrand and Heroux, 1987;

99 Bertrand, 1990; Bertrand, 1993; Yang and Hesse, 1993; Bertrand and Malo, 2001; Bertrand et al., 2003),

100 and Conodont Alteration Index (CAI; Goodarzi and Norford, 1985; Hoffknecht, 1991; Gentzis et al.,

101 1996). Moreover, the graptolite reflectance has also been successfully correlated with some inorganic

102 paleothermal indices, including the illite “crystallinity” index (e.g. Kemp et al., 1985; Oliver, 1988;

103 Hoffknecht, 1991; Rantitsch, 1995, 1997; Suchý et al., 2015), and metamorphic index minerals in pre-

104 greenschist meta-sedimentary rocks (e.g. Malinconico, 1992). Graptolite reflectance is commonly related

105 to the lithology, and it will be higher in shales than in (Link et al., 1990). Redox conditions

106 may have an impact on graptolite reflectance, and higher reflectance was observed in graptolites deposited

107 under oxic environments in comparison with those from anoxic environments (Cole, 1994). Weathering

108 may affect graptolite reflectance (Goodarzi and Norford, 1985; Hoffknecht, 1991; Goodarzi et al., 1992a).

109 Raman spectroscopy of graptolites was widely studied and used to assess organic maturity (Suchý et al.,

110 2004; Liu et al., 2013; İnan et al., 2016; Mumm and İnan, 2016; Morga and Pawlyta, 2018; Wang et al.,

111 2019a). However, these studies were mostly focused on a small number of samples representing a very

112 limited range of thermal maturity. Thus, the physical and chemical characteristics of graptolites in wider

113 thermal maturity settings is reviewed herein for Lower Paleozoic sediments from a number of localities

114 based on the literature and new unpublished data is presented. In reviewing the optical characteristics,

115 chemical composition, and microstructure of the graptolites, their implications for the global petroleum

116 industry were also discussed.

117 2. The biological structure and composition of graptolites

118 Graptolites are extinct colonial planktonic hemichordate invertebrates that lived mainly in the early 119 Paleozoic ocean (Clarkson, 1981). Their structure has been widely studied and is illustrated in Fig. 1

120 (Moore, 1955; Clarkson, 1981; Crowther, 1981). The rhabdosome (colony) may have more than one

121 branch (stipe) (Fig. 1a). Along a stipe, thecae house the individual zooids (Fig. 1b, 1c), which were

122 connected by a stolon to other individuals through the common canal (Fig. 1c). The periderm material of

123 the rhabdosome is generally composed of two layers, fusellar tissue and cortical tissue (Fig. 1e)

124 (Crowther, 1981). The periderm with fusellar layers, thecae and common canal may be observed in

125 reflected light under the optical microscope as discussed below.

126 Initially, studies of the chemical structure of graptolites were related to the periderm based on the

127 textures and phylogeny, and these studies assumed that graptolite periderm is chitinous (Eisenack, 1932;

128 Kozlowski, 1949). Later, a collagen-like structure for the cortical and fusellar tissues was proposed for

129 graptolites (Fig. 1) (Towe and Urbanek, 1972; Crowther, 1981; Bates and Kirk, 1986; Urbanek and

130 Mierzejewski, 1986; Liu et al., 1996). Both collagen and chitin are supportive tissues, which mainly serve

131 to support cellular structures (Tasch, 1980). Collagen is a glycoprotein, and chitin with a molecular

132 formula of C32H54O21N4 is a nitrogenous carbohydrate forming a N-acetyl glucosamine groups polymer

133 (Leninger, 1975; Bustin et al., 1989). More recent studies indicate, however, that the residual graptolite

134 periderms are composed of complex aliphatic polymers that are immune to base hydrolysis rather than

135 collagen (e.g. Briggs et al., 1995; Gupta et al., 2006). Aliphatic components of graptolite periderms are

136 probably derived from the graptolite itself via in situ polymerization (e.g. Gupta et al., 2006).

137 138

139 Fig. 1. The biological structure of graptolites showing their theca, aperture, common canal, virgella and periderm (Modified after Moore,

140 1955; Clarkson, 1981). A) Rhabdosome (colony) with two stipes (branches); B) close-up of rhabdosome structures; C) Close-up of two

141 thecae which house individual zooids; D) transverse cross-section of rhabdosome; E) periderm structure.

142

143 3. Sampling and methods 144 The graptolite-bearing sediments examined for this study included the Wufeng–Longmaxi

145 Formations, Pingliang Formation, China (Luo et al., 2017; Luo et al., 2018; Wang et al., 2019a); Liteň

146 Formation from Czech (Suchý et al., 2002); and Alum Shales from Estonia and Sweden (Petersen et al.,

147 2013; Sanei et al., 2014; Luo et al., 2018). Detailed geological information can be found in these above

148 references. Two graptolite-bearing Wufeng–Longmaxi samples (sample ID: CKMB-Y1 (GRor = 1.32%)

149 and CKMB-Y2 CKMB-Y1 (GRor = 1.28%)) (Wang et al., 2019a) and an upper coal (VRor

150 = 1.07%) were selected to conduct artificial maturation at 350℃, 400℃, 450℃ and 500℃ for three days,

151 respectively. The natural and heated samples were cut parallel or perpendicular to bedding in order to

152 make polished sections on EcoMet 250 with AutoMet 250 for the maceral observation and reflectance

153 measurement. Organic petrography and reflectance were done on a Leica 4500P microscope with 154 CRAIC/MPS 200 microscope photometer. The maximum-minimum and random reflectances were

155 measured under polarized and non-polarized light, respectively, as described in Luo et al. (2016); Luo et

156 al. (2017); Luo et al. (2018).

157 For scanning electron microscopy (SEM), sediment was cut, polished using emery paper, and milled

158 using Ar ion milling. For study of pores, graptolites were found and then marked using a diamond lens

159 under an optical microscope. The SEM study was on a Zeiss Crossbeam 540 Focused Ion Beam-SEM

160 (FIB-SEM).

161 4. Optical characteristics of graptolite periderm

162 Kurylowicz et al. (1976) reported the reflectance characteristics of graptolites and proposed that their

163 optical properties are similar to vitrinite. Clausen and Teichmüller (1982) described the morphology and

164 reflectance of graptolites from Germany and Sweden, with maximum reflectances ranging from 0.8% to

165 10.0%. In 1980–1990s, Goodarzi and colleagues extensively studied the optical properties of Middle

166 to Upper graptolites from Poland, Sweden, Turkey and Canada (Goodarzi, 1984,

167 1985a; Goodarzi and Norford, 1985; Goodarzi et al., 1985; Goodarzi and Norford, 1987; Bustin et al.,

168 1989; Goodarzi and Norford, 1989; Riediger et al., 1989; Link et al., 1990; Goodarzi et al., 1992a;

169 Goodarzi et al., 1992b; Gentzis et al., 1996), which advanced the understanding of the petrographic

170 characteristics of graptolites. In recent years, graptolite-bearing sediments have become significant targets

171 of shale gas exploration, and a resurgence of work has been conducted on the organic petrology of

172 graptolites (Petersen et al., 2013; Sanei et al., 2014; Haeri-Ardakani et al., 2015; Caricchi et al., 2016;

173 İnan et al., 2016; Lavoie et al., 2016; Luo et al., 2016; Ma et al., 2016; Mumm and İnan, 2016; Cheshire

174 et al., 2017; Luo et al., 2017; Cardott and Curtis, 2018; Luo et al., 2018; Morga and Kamińska, 2018;

175 Morga and Pawlyta, 2018; Reyes et al., 2018; Synnott et al., 2018; Wang et al., 2019a).

176 Microscopically, graptolites can be identified by their morphology, e.g., fusellar layer, granularity,

177 other visible structures as well as anisotropy (Kurylowicz et al., 1976; Teichmüller, 1978; Clausen and

178 Teichmüller, 1982; Goodarzi, 1984, 1985a; Bertrand and Heroux, 1987; Goodarzi and Norford, 1987,

179 1989; Bustin et al., 1989; Riediger et al., 1989; Bertrand, 1990; Petersen et al., 2013; Luo et al., 2016;

180 Luo et al., 2017; Luo et al., 2018). Two textures are apparent in graptolites, i.e., non-granular (NGG) and

181 granular (GG), in Paleozoic sediments worldwide, e.g., Poland, Sweden, Turkey and Canada (Goodarzi,

182 1984, 1985a; Goodarzi and Norford, 1985; Goodarzi et al., 1985; Goodarzi and Norford, 1987, 1989;

183 Goodarzi et al., 1992a; Goodarzi et al., 1992b), the Alum Shale from Europe (Petersen et al., 2013; Sanei 184 et al., 2014; Luo et al., 2018), and the Wufeng–Longmaxi Formations and Pingliang Formation from

185 China (Fig. 2) (Wang et al., 1993; Luo et al., 2016; Luo et al., 2017; Luo et al., 2018; Wang et al., 2019a).

186 In general, GG and NGG are found in carbonates and mudstones, respectively (Goodarzi and Norford,

187 1985; Goodarzi and Norford, 1987; Petersen et al., 2013). NGG and GG were thought to be derived from

188 different sections of the rhabdosome. While the former may be part of the wall with fusellar layers, the

189 latter may be derived from the common canal (Goodarzi, 1984). However, according to recent artificial

190 maturation studies on Estonian Alum Shale with GG, this texture is completely altered to NGG when

191 GRor reached up to 1.24% after heating for 3 days at 400 ℃. This indicates that pristine GG may be altered

192 to NGG due to thermal stress (Luo et al., 2018).

193 GG has weaker anisotropy and lower reflectance than NGG (Figs. 2b and 3) (Goodarzi, 1984;

194 Goodarzi and Norford, 1987; Petersen et al., 2013; Luo et al., 2016; Luo et al., 2017; Luo et al., 2018).

195 Suchý et al. (2002) found that NGG random reflectance has a positive correlation with GG random

196 reflectance in Lower Silurian sediments from the Barrandian Basin, Czech Republic. In rotation of the

197 objective stage under polarized light, NGG will display extinction twice, and thus two GRomax and two

198 GRomin can be measured (Figs. 3a, b and 4) (Luo et al., 2016; Luo et al., 2018), however, GG does not

199 show such clear patterns (Fig. 3c and d). 200

201 Fig. 2 Granular-textured graptolites (GG) and non-granular-textured graptolites (NGG) from Chinese and Estonian sediments. (a) GG in

202 low-maturity Alum Shale from Estonia (GRor = 0.65%), non-polarized light; (b) GG and NGG in high-maturity shale from Chongqing,

203 China (GRomax = 2.57%), non-polarized light (Luo et al., 2018); (c) grey black NGG in low-maturity shale from Gansu, China (GRor =

204 0.48%), non-polarized light; (d) NGG in overmature shale from Chongqing, China (GRor = 3.85%), showing strong anisotropy, polarized

205 light; (e) NGG in overmature shale from Chongqing, China (GRor = 3.77%), showing strong anisotropy, polarized light; (f) NGG in

206 overmature shale from Chongqing, China (GRomax = 4.47%), showing fusellar layers, polarized light (modified from Luo et al., 2018).

207 (a-e) sections perpendicular to bedding; (f) sections parallel to bedding.

208 5 6 (a) (b) 5 4

4

3 (%)

(%) 3

o o R R 2 2

1 1

0 0 0 40 80 120 160 200 240 280 320 360 0 40 80 120 160 200 240 280 320 360 209 Degree Degree

2.5 2.2 (c) (d)

2.1

2.2 (%)

(%) 2

o o

R R 1.9 1.9

1.6 1.8 0 40 80 120 160 200 240 280 320 360 0 40 80 120 160 200 240 280 320 360 210 Degree Degree

211 Fig. 3. Reflectance measured on NGG (a and b) and GG (c and d) under polarized light in a circular rotation of the microscope stage.

212

213 Fig. 4. Photomicrographs displaying extinction characteristics of the same NGG from the Wufeng–Longmaxi shales in sections

214 perpendicular to bedding with 360° rotation of the microscope stage, oil immersion, polarized light. The angle between each picture is 215 90 degrees. The graptolite reflectance (%) is given beside the measurement site (red square).

216

217 In general, NGG is more frequently found than GG in most Paleozoic sediments (Goodarzi and

218 Norford, 1987; Petersen et al., 2013; Luo et al., 2016), thus, only NGG will be discussed in the main

219 discussion of this paper. In polished blocks, low-maturity graptolites are often similar to natural bitumen

220 stringers, but they are more angular than bitumen and often show typical graptolitic features (Figs. 2, 4

221 and 5), and they display a plate-like appearance at high maturity (Fig. 2). NGG is blocky in sections

222 parallel to bedding, but may display a long stipe-like morphology in sections normal to bedding (Figs. 2

223 and 5) (Link et al., 1990). The fusellar layers are observed in sections parallel to bedding (Fig. 2). NGG

224 was divided into two classes, lath- and blocky-shaped (Riediger et al., 1989). Blocky-shaped graptolites

225 show lower maximum reflectance and bireflectance (BRo) than lath-shaped graptolites. Some NGG of the

226 Liteň Formation from Czech Republic displays weak brown fluorescence (Fig. 5), and Hoffknecht (1991)

227 and Wang et al. (2019a) also found similar phenomena in low-maturity graptolites. 228

229 Fig. 5. NGG from the Czech Republic (Liteň Formation, Barrandian area) with weak brown fluorescence of the Liteň Formation in

230 sections perpendicular to bedding. (a, c, e) under reflected light; (b, d, f) the same field as (a, c, e), under fluorescence light. (a-b) GRor =

231 0.85%; (c-f) GRor = 0.83%.

232 233 5. Physical and chemical properties of graptolites as thermal maturity indicators 234 5.1 Chemical structure changes of graptolites with maturation 235 OM is sensitive to pressure and temperature (Khavari-Khorasani, 1975; Teichmüller, 1982; Goodarzi

236 and Norford, 1985). Optical properties of OM include reflectances, refractive and absorptive index (van

237 Krevelen, 1961). Reflectance of OM is related to the aromatic structure of organic molecules; reflectance 238 and aromaticity increase continuously with increase in maturity. The trends of optical properties of OM

239 over the visible spectrum (400–700nm) are used to determine molecular structural changes occurring

240 during maturation (Khavari-Khorasani, 1975; Goodarzi, 1985a; Goodarzi and Macqueen, 1990).

241 Structurally, OM is composed of ordered and amorphous carbons (Cartz and Hirsch, 1960; van

242 Krevelen, 1961; Goodarzi, 1985a). The ratio of ordered (aromatic) carbons to amorphous (alicyclic side

243 chains, aliphatic) carbons increases with maturation/heat treatment (Cartz and Hirsch, 1960; van Krevelen,

244 1961) in two phases:

245 ① Devolatilization of amorphous carbon increase in aromaticity and size of the polynuclear systems

246 during early stage of maturation and low temperature heat treatment;

247 ② Transformation of 2D (turbostratic) to 3D (graphitic) ordering, which occurs at high maturity

248 and at temperature >600 ℃, results in increase and better ordering of the aromatic structure (Cartz

249 and Hirsch, 1960; van Krevelen, 1961; Goodarzi and Murchison, 1972; Goodarzi, 1985a).

250 Dispersion of optical properties of graptolite periderm is similar to that of bitumen and vitrinite in

251 the visible spectrum and three spectral patterns are documented:

252 ① Low-maturity (CAI = 1) graptolite behaves similarly to low rank vitrinite and bitumen with

253 curves of all optical parameters decreasing with increasing wavelength from blue to red (Figs. 6 and

254 7), indicating low content of aromatic carbon (Marshall and Murchison, 1971; Khavari-Khorasani,

255 1975);

256 ② Moderately-mature (CAI = 4) graptolite behaves similarly to semi-anthracite and impsonite

257 bitumen. The optical parameters are nearly flat, from blue to red (Figs. 6 and 7), indicating gradual

258 molecular variations, e.g., greater condensed aromatic carbon (Marshall and Murchison, 1971;

259 Khavari-Khorasani, 1975);

260 ③ Highly mature graptolite (CAI = 3.5–5) behaves similarly to anthracite and pyrobitumens. The

261 trends of optical indices rise continuously from blue to red (Figs. 6 and 7), similar to the pattern from

262 a highly aromatic, condensed molecular structure (Marshall and Murchison, 1971; Cook et al., 1972;

263 Khavari-Khorasani, 1975; Goodarzi, 1985a). Graptolites at this stage develop macro-properties

264 similar to graphite (Teichmüller et al., 1979). 265

266 Fig. 6. Dispersion of maximum reflectance (a) in air (%Ramax) and (b) in oil (%Romax) of graptolite fragments with increased maturity

267 based on their CAI (conodont alteration index), shown as number beside each spectrum for graptolite in (L), argillaceous

268 limestone (A) and shale (S) matrices (after Goodarzi, 1985a).

269

270 Fig. 7. Dispersion of refractive (a) and (b) absorptive indices of graptolite fragments with increased maturity based on their CAI, shown

271 as number beside each spectrum for graptolites in limestone (L), argillaceous limestone (A) and shale (S) matrices (after Goodarzi,

272 1985a). 273

274 5.2 Graptolite reflectance 275 5.2.1 Graptolite maximum, minimum and random reflectance

276 Similar to vitrinite (Davis, 1978), GRomax and graptolite minimum reflectance (GRomin) are observed

277 in sections parallel and normal to bedding, respectively (Goodarzi, 1984; Link et al., 1990; Goodarzi et

278 al., 1992a). However, in some samples, GRomax in sections parallel to bedding is lower than that in sections

279 normal to bedding (Table 1 and Fig. 8), which may be because the graptolites in these samples are not

280 truly parallel to bedding or normal to bedding due to an unusual burial process and/or sample preparation

281 (Malinconico, 1992, 1993; Luo et al., 2017). In sections parallel to bedding, GRomax positively correlates

282 with GRomin (Fig. 9), whereas in sections perpendicular to bedding, such relationship cannot be observed

283 (Hoffknecht, 1991; Luo et al., 2017). This is because GRomin in sections parallel to bedding is generally

284 equal to the graptolite intermediate reflectance (GRoint) (Luo et al., 2017).

285

286 Table 1. Comparison of graptolites reflectance (including maximum, minimum and bireflectance) in sections parallel and normal to

287 bedding. Part of the data for the Wufeng–Longmaxi graptolites are derived from Luo et al. (2017), and the data of the Upper Silurian

288 graptolites of Turkey are from Goodarzi (1984).

Sections parallel to bedding Sections normal to bedding Depth Sample no. Location Formation Lithology GRomax GRomin BRo GRomax GRomin BRo (m) (%) (%) (%) (%) (%) (%)

DTB-6-9 Datianba, China Outcrop Shale 3.41 1.92 1.49 3.68 0.99 2.69 Wufeng- JX-4 Jinxi, China Outcrop Shale 5.48 2.61 2.87 5.61 0.64 4.98 Longmaxi QQ1-3 Qianjiang, China 794.00 Shale 4.76 3.14 1.62 5.64 0.75 4.90 Upper No. 2 Tufanbeyli, Turkey Outcrop Shale 4.61 4.17 0.44 4.42 1.08 3.34 Silurian

289

290 Fig. 10 displays a comparison of GRor in sections parallel and perpendicular to bedding. In contrast

291 to GRomax, GRor in sections perpendicular to bedding, especially in samples with higher thermal maturity,

292 is much lower than that in sections parallel to bedding, similar to the results in the Qusaiba Hot Shales

293 from Saudi Arabia (İnan et al., 2016), and is much more concentrated than that in sections parallel to

294 bedding as indicated by their lower SD values (Fig. 10). In general, unimodal histograms of GRomax and

295 GRor were found in investigated samples, especially in sections perpendicular to bedding (Figs. 8 and 10)

296 (Cole, 1994; Haeri-Ardakani et al., 2015; Lavoie et al., 2016; Luo et al., 2018; Reyes et al., 2018). In 297 addition, NGG reflectance in carbonate rocks is lower than that of shales at similar depths, which may be

298 due to oxidation caused by carbonate dissolution (Goodarzi and Norford, 1985; Link et al., 1990). This is

299 similar to that of VRor in different lithologies (Goodarzi et al., 1988; Goodarzi et al., 1993).

300

301

302 Fig. 8. Comparison of GRomax of the Wufeng–Longmaxi sediments in (a, c and e) sections parallel to bedding and (b, d and f) sections

303 perpendicular to bedding from Chongqing.

304 305

306 Fig. 9. The linear coalification trend for graptolites based on the plot of GRomax VS. GRomin (after Goodarzi, 1984).

307

308

309 Fig. 10. Comparison of GRor of the Wufeng–Longmaxi sediments in (a, c, e) sections parallel to bedding and (b, d, f) sections

310 perpendicular to bedding from Chongqing (cited from Luo et al., 2019). 311

312 From the methodological point of view, the main difference between the measurement technique of

313 vitrinite and graptolite reflectance has to do with the sample preparation. Graptolite maturity

314 determination was generally on the basis of whole-rock polished blocks in sections parallel to bedding

315 and using of GRomax (Goodarzi, 1984; Goodarzi and Norford, 1985; Goodarzi and Norford, 1987, 1989;

316 Gentzis et al., 1996). GRomax will increase with burial depth and is a useful tool to unravel the thermal

317 maturity of Lower Paleozoic deposits (Clausen and Teichmüller, 1982; Goodarzi, 1984, 1985a; Goodarzi

318 and Norford, 1987, 1989; Riediger et al., 1989; Link et al., 1990; Gentzis et al., 1996; Luo et al., 2016;

319 Luo et al., 2017). However, other researchers have used GRor to assess maturity levels (Bertrand and

320 Heroux, 1987; Bertrand, 1990; Bertrand, 1993; Cole, 1994; Bertrand and Malo, 2001; Bertrand et al.,

321 2003; Petersen et al., 2013; Sanei et al., 2014; İnan et al., 2016; Yang, 2016; Luo et al., 2018; Reyes et al.,

322 2018; Synnott et al., 2018; Wang et al., 2019a).

323 Based on data from published literature, GRomax shows a strong positive correlation with GRor in

324 sections perpendicular to bedding, and the former is around twice as large as the latter (Fig. 11). The

325 comparison of SD for GRomax and GRor is shown in Fig. 12. In general, SD is greater for GRomax than that

326 of GRor. Therefore, for routine maturation study, measurements of GRor is as accurate and as useful as

327 those of GRomax. 328

6 Wang et al., 2019a y = 0.52x + 0.43 Luo et al., 2018 r = 0.95 Malinconico, 1993 4 Link et al., 1990

2 Mean random Mean random reflectance graptoliteof(%)

0 0 3 6 9 Mean maximum reflectance of graptolite (%) 329

330 Fig. 11. Mean GRomax vs. mean GRor in sections perpendicular to bedding (Link et al., 1990; Malinconico, 1993; Luo et al., 2018; Wang

331 et al., 2019a).

332 0.9 0.9 (a) Malinconico, 1992 (b) Luo et al., 2018 Luo et al., 2018 Petersen et al., 2013 Wang et al., 2019a Bertrand et al., 2003 Lavoie et al., 2016 0.6 0.6 Williams et al. 1998 Wang et al., 2019a y = 0.03x + 0.07

r = 0.51

SD SD (%) SD SD (%)

0.3 0.3

y = 0.07x + 0.07 r = 0.75 0 0 0 3 6 9 0 2 4 6 333 Mean maximum reflectance of graptolite (%) Mean random reflectance of graptolite (%)

334 Fig. 12. The relationship between (a) mean GRomax in sections perpendicular to bedding, (b) mean GRor and SD (Malinconico, 1992;

335 Williams et al., 1998; Bertrand et al., 2003; Petersen et al., 2013; Lavoie et al., 2016; Luo et al., 2018; Wang et al., 2019a).

336

337 5.2.2 Graptolite bireflectance (BRo)

338 BRo in sections parallel to bedding is much lower than that in sections perpendicular to bedding,

339 especially in overmature sediments (Table 1) (Goodarzi, 1984). The anisotropy of graptolites will increase

340 with increasing thermal maturity as suggested by positive correlation between GRomax and BRo (Luo et

341 al., 2017). Vitrinite displays lower bireflectance than that of graptolite at similar maturation levels because

342 of the biaxial negative property of graptolites (Bustin et al., 1989; Hoffknecht, 1991; Luo et al., 2017).

343 5.2.3 The coalification path of the graptolites

344 The plot of GRomax vs. GRomin can be used to assess the coalification path of the graptolites (Fig. 9).

345 However, it should be noted that GRomin in the Fig. 9 was achieved in section parallel to bedding, and is

346 nearly equal to GRoint (Goodarzi, 1984; Goodarzi and Norford, 1985; Goodarzi and Norford, 1987;

347 Goodarzi et al., 1992a; Malinconico, 1993; Luo et al., 2017; Luo et al., 2018). When all the graptolite

348 reflectance data from the published literature are plotted in Fig. 9, a linear coalification trend can be

349 determined for the graptolites measured parallel to bedding (Fig. 9) (Goodarzi, 1984; Goodarzi and

350 Norford, 1985; Goodarzi and Norford, 1987; Goodarzi et al., 1992a; Malinconico, 1993; Luo et al., 2017;

351 Luo et al., 2018).

352 5.2.4 The relationship between graptolite reflectance and other paleothermal indices

353 5.2.4.1 Comparison with other organoclasts reflectance

354 Similar to vitrinite and solid bitumen (see also Ferreiro Mählmann and Le Bayon, 2016 for a recent 355 review on vitrinite and solid bitumen thermal maturity studies), graptolite reflectance displays an

356 increasing trend with the increasing burial depth, and can determine organic maturity in Lower Paleozoic

357 sediments without vitrinite, although no consensus of the exact relationship between these indices has yet

358 been reached (Table 2 and Fig. 13) (Clausen and Teichmüller, 1982; Goodarzi, 1984, 1985a, b; Bertrand

359 and Heroux, 1987; Goodarzi and Norford, 1987; Teichmüller, 1987; Bustin et al., 1989; Goodarzi and

360 Norford, 1989; Bertrand, 1990; Goodarzi et al., 1992a; Bertrand, 1993; Cole, 1994; Gentzis et al., 1996;

361 Bertrand and Malo, 2001; Bertrand et al., 2003; Petersen et al., 2013; Luo et al., 2018; Synnott et al.,

362 2018). Goodarzi (1985a) found that the optical dispersion of graptolites is similar to vitrinite and bitumen,

363 indicating that their physical and chemical nature is sensitive to temperature and can be used as a proxy

364 for organic maturity in Lower Paleozoic deposits. The graptolites show higher reflectance and stronger

365 anisotropy than that of solid bitumen, chitinozoans and scolecodonts (Goodarzi, 1984, 1985b; Goodarzi

366 and Norford, 1987, Bertrand et al., 2003; Petersen et al., 2013). According to Yang and Hesse (1993), the

367 bireflectance of graptolites displays a similar trend to that of pyrobitumen when Ro < 4.5–5.0 %. Other

368 researchers suggested that graptolite reflectance is similar to that of chitinozoan for both natural and

369 artificially matured samples (Bertrand and Heroux, 1987; Bertrand, 1990; Cole, 1994; Reyes et al., 2018).

370 Zooclast reflectance (including chitinozoan, graptolite, and scolecodonts) was also compared with

371 vitrinite and solid bitumen reflectances in Lower Paleozoic deposits from Canada (Bertrand and Heroux,

372 1987; Bertrand, 1990; Bertrand, 1993; Bertrand and Malo, 2001; Bertrand et al., 2003). In these Canadian

373 studies, graptolite reflectance was slightly lower than, or similar to, vitrinite reflectance (Bertrand, 1990)

374 (Table 2 and Fig. 13). In a similar way, Yang and Hesse (1993) established that GRor is lower than VRor

375 when GRor < 2.6%, but graptolite has higher reflectance than vitrinite when GRor > 2.6%. The Silurian

376 Qusaiba Shale from Saudi Arabia deposited under anoxic environments has lower graptolite random

377 reflectance than that deposited in oxic environments (Cole, 1994). Zhong and Qin (1995) proposed a

378 formula between EqVRo and GRor based on vitrinite-like particle reflectance (Rvl) and GRor in the

379 sediments from Tarim Basin and South China. In recent years, various new equations have been proposed

380 to illustrate the relationship between EqVRo and GRor or GRomax (Petersen et al., 2013; Colţoi et

381 al., 2016; Synnott et al., 2018; Luo et al., 2018; Wang et al., 2019a) (Table 2 and Fig. 13).

382 The equations proposed by Cole (1994) (anoxic); Petersen el al., (2013) and Coltoi et al., (2016) are very

383 close to each other; whereas the formulas presented by Bertrand (1990) and Luo et al. (2018) are

384 essentially coincident and close to the 1:1 line (Fig. 13). 385 Table 2. Various equations describing the relationship between EqVRo and GRor or GRomax.

GRor data Equations Source Formation range

Log10 GRor = -0.04+1.10×Log10 EqVRo (Bertrand, 1990) 0.5–3.0% Upper Gaspe Limestone Group and Chaleurs Group, Canada

EqVRo = 0.8 ×GRor (anoxic) (Cole, 1994) 0.62–2.05% Silurian Qusaiba Shale, Saudi Arabia

EqVRo = 0.65 ×GRor (oxic)

EqVRo = 0.882 ×GRor-0.366 (Zhong and Qin, 1.8–4.9% -Ordovician, China 1995)

EqVRo = 0.73 ×GRor + 0.16 (Petersen et al., 0.47–2.14% Alum Shale, Scandinavia 2013)

EqVRo = 0.785×GRor + 0.05 (Colţoi et al., 2016) No data Silurian intervals, Romania

EqVRo = 0.232+ 0.499×GRor (Synnott et al., 0.59–1.02% Cape Phillips Formation, Canada 2018)

EqVRo = 1.055×GRor-0.053 (Luo et al., 2018) 0.65–4.03% Wufeng–Longmaxi Formation

EqVRo = 0.546×GRomax +0.35

EqVRo = 0.97×GRor-0.2 (2.19% < (Wang et al., 2019a) 1.21–4.91% Wufeng–Longmaxi Formation

GRor < 3.5%);

EqVRo = 0.22×GRor +2.55(GRor > 3.5%)

EqVRo = 0.99×GRor + 0.08; This study

EqVRo = 0.515×GRomax + 0.506

386

6 Bertrand (1990)

Cole (1994) (anoxic)

Cole (1994) (oxic)

Zhong and Qin (1995)

4 Petersen et al. (2013)

Colţoi et al. (2016)

(%) Synnott et al. (2018) o

Luo et al. (2018) (GRo) EqvR Luo et al. (2018) (GRomax) 2 Wang et al. (2019a) (2.19% < GRor < 3.5%)

Wang et al. (2019a) (3.5% < GRor < 4.91%)

This study (GRo)

This study (GRomax)

1:1 line 0 0 2 4 6 387 GRo (%) (except where noted)

388 Fig. 13. The various relationships between GRo or GRomax and EqvRo.

389 390 In China, graptolites were usually interpreted as solid bitumen or vitrinite-like particles in the

391 Wufeng–Longmaxi Formations, due to the misidentification on OM, leading to a long-term debate on

392 their thermal maturity. Luo et al. (2018) have described how to discriminate between graptolite and

393 bitumen in those formations. NGG displays a smoother surface, stronger anisotropy, and higher random

394 and maximum reflectance than solid bitumen based on observation and measurement under polarized and

395 non-polarized light (Luo et al., 2018; Wang et al., 2019a). In the Wufeng–Longmaxi Formations, the solid

396 bitumen random reflectance (SBor) has strong positive correlations with GRor and GRomax in sections

397 perpendicular to bedding (Fig. 14a and b) (Luo et al., 2018; Wang et al., 2019a). Data from other

398 graptolite-bearing sediments (Yang and Hesse, 1993; Bertrand et al., 2003; Yang, 2016; Reyes et al., 2018)

399 are plotted with the Wufeng–Longmaxi data in Fig. 14c and d, which indicates that SBor and GRor are

400 nearly equivalent, very different from their relationship in Fig. 14a. This is because the reflectances from

401 Yang and Hesse (1993), Bertrand et al. (2003), Yang, (2016), and Reyes et al. (2018) were measured in

402 random orientation whereas the data in the two Wufeng-Longmaxi studies were measured on sections

403 perpendicular to bedding (Luo et al., 2018; Wang et al., 2019a).

9 12 (a) (b) Wang et al., 2019a

Luo et al., 2018 y = 2.12x - 0.57 r = 0.92 6 8

3 4

y = 1.19x - 0.09 Wang et al., 2019a

r = 0.95 (%) graptolite reflectance of maximum Mean

Mean random Mean random reflectance graptoliteof(%) Luo et al., 2018

0 0 0 2 4 6 0 2 4 6 Mean random reflectance of solid bitumen (%) 404 Mean random reflectance of solid bitumen (%) 9 9 (c) Wang et al., 2019a (d) Reyes et al., 2018 Luo et al., 2018 Yang et al., 2016 Reyes et al., 2018 outlier outlier Bertrand et al., 2003 Yang et al., 2016 Yang and Hesse, 1993 6 Bertrand et al., 2003 6 Yang and Hesse, 1993

3 3

y = 1.05x + 0.20 y = 0.98x + 0.32

r = 0.94 Mean random Mean random reflectance graptoliteof(%) Mean random Mean random reflectance graptoliteof(%) r = 0.95

0 0 0 2 4 6 0 2 4 6 Mean random reflectance of solid bitumen (%) Mean random reflectance of solid bitumen (%) 405

406 Fig. 14. The relationship between mean SBor and (a, c and d) mean GRor, and (b) mean GRomax (Yang and Hesse, 1993; Bertrand et al.,

407 2003; Yang, 2016; Luo et al., 2018; Reyes et al., 2018; Wang et al., 2019a).

408

409 5.2.4.2 Comparison with CAI

410 Some studies have correlated the relationship between GRomax and VRor through CAI. When

411 sediments have GRomax of 0.6–1.2%, it means that they are immature (~0.2–0.5% VRor, CAI 1.5); when

412 GRomax is 1.2–2.2%, sediments are in the oil window (~0.5–1.30% VRor, CAI: 1.5–2.5) (Fig. 15)

413 (Goodarzi and Norford, 1989; Goodarzi et al., 1992a; Gentzis et al., 1996). However, these results should

414 be used very carefully, because a CAI generally implies a range of thermal maturity (Epstein et al., 1977).

415 Graptolite and vitrinite reflectances are more sensitive to thermal maturity than CAI, especially in

416 overmature sediments (Goodarzi and Norford, 1985). While graptolite reflectance displays an excellent

417 correlation with CAI at low thermal maturity (CAI: 1–4), their correlation is not clear in sediments with

418 higher CAI (Goodarzi and Norford, 1985; Goodarzi et al., 1992a). 419

420 Fig. 15. The relationship between CAI and GRomax (Goodarzi and Norford, 1989; Goodarzi et al., 1992a; Gentzis et al., 1996).

421

422 5.2.4.3 Comparison with Tmax

423 Graptolite reflectance has also been compared to Tmax for the same samples (Cole, 1994; Petersen et

424 al., 2013; İnan et al., 2016; Synnott et al., 2018). Cole (1994) studied organic maturity of Silurian Qusaiba

425 deposits, Saudi Arabia, using zooclasts reflectance and conventional maturity indicators (e.g., Tmax), and

426 proposed that GRor will increase with the increasing oxygen contents of the depositional environments.

427 He found that EqVRo of the sediments deposited under anoxic environments is equal to 80% of GRor,

428 whereas the EqVRo of the sediments deposited under oxic environments is equal to 65% of GRor (Table

429 2) (Cole, 1994). Petersen et al. (2013) have determined the relationship between GRor and EqVRo on the

430 basis of Tmax, which can be expressed as: EqVRo = 0.73×GRor + 0.16 (Table 2). It should be noted that

431 GRor displays a bimodal distribution due to the random orientation of the polished blocks in their study.

432 İnan et al. (2016) determined the thermal maturity of the hot shales from Saudi Arabia based on graptolite

433 reflectance, organic geochemical and spectroscopic methods. Other studies on Tmax and graptolite

434 reflectance are conflicting and inconclusive (Table 2) (e.g., Petersen et al., 2013; Synnott et al., 2018).

435 The difficulty in using Tmax is mostly due to unreliable values of Tmax, the influence of kerogen type, and

436 low S2 values (e.g., Petersen et al., 2013; Synnott et al., 2018). 437 5.2.4.4 Comparison of artificial thermal-treatment of graptolites and coals/vitrinite

438 Laboratory experiments where graptolites and vitrinite with similar maturity were heat treated (220–

439 600 ℃) under similar conditions have shown that graptolites develop high reflectance at 600 ℃ (Bustin

440 et al., 1989). Reyes et al. (2018) used hydrous pyrolysis on immature graptolite-bearing sediments

441 (GRor=0.55%) of the Boas River Formation treated to temperatures of 310 to 350℃, and proposed the

442 relationship between GRor and Rvl as the following equation: Rvl = 0.79×GRor. In contrast, Luo et al. (2018)

443 heated graptolite-bearing sediment, with wider maturity (GRor = 0.65 to 4.03%), through a temperature

444 range of 350 ℃ and 550 ℃ and found that their reflectance is similar to that of the vitrinite.

445 Comparison of all published data and data from this study on heat–treated and natural graptolites

446 and vitrinite (or vitrinite-like particles) (Bustin et al., 1989; Bertrand, 1990; Luo et al., 2018; Reyes et al.,

447 2018), regardless of the difference of the experimental methods, indicates a strong and positive

448 relationship (Fig. 16). This relationship can be expressed as:

449 EqVRo = 0.99×GRor + 0.08 (1)

450 This equation is essentially coincident with Bertrand (1990) and Luo et al. (2018), and close to the

451 1:1 line (Fig. 9).

452 Equation (1) combined with Fig. 11 can correlate GRomax to that of EqVRo and it can be expressed

453 as:

454 EqVRo = 0.515×GRomax + 0.506 (2)

455

6

y = 0.99x + 0.08 r = 0.99

4

Luo et al., 2018 2 Bertrand, 1990

Vitrinite random Vitrinite random reflectance (%) Bustin et al., 1989 Reyes et al., 2018 This study 0 0 2 4 6 Graptolite random reflectance (%) 456

457 Fig. 16. The relationship between VRor and GRor. The Rvl (reflectance of vitrinite-like particles) values adopted from Reyes et al. (2018) 458 were converted to EqVRo according to the equation proposed by Xiao et al. (2000).

459

460 5.3. The chemical composition of graptolite periderm 461 The electron microprobe, FTIR and Raman spectrum have been widely used to study the chemical

462 composition of various organic macerals including graptolite fragments (Zerda et al., 1981; Green et al.,

463 1983; Jehlička and Bény, 1992; Wopenka and Pasteris, 1993; Bustin et al., 1993; Mastalerz and Bustin,

464 1993b, a, 1995, 1996, 1997; Ward and Gurba, 1999; Kelemen and Fang, 2001; Ward et al., 2005, 2007,

465 2008; Chen et al., 2012; Chen et al., 2014; Wang et al., 2014; Wilkins et al., 2014; Wilkins et al., 2015;

466 Chen et al., 2015b; Lünsdorf, 2016; Cheshire et al., 2017; Henry et al., 2018).

467 5.3.1 Electron microprobe

468 Recently, the electron microprobe has been used to identify elemental compositions of graptolites in

469 the Silurian Llandovery–Ludlow shales from northern Poland (Morga and Kamińska, 2018). GRor of these

470 shales ranges between 1.30% and 1.83%. Carbon is the predominant element in graptolite periderm;

471 carbon, oxygen, nitrogen and sulfur contents range from 84.92 to 91.45 %, 2.56 to 8.43 %, 1.43 to 2.89 %

472 and 0.02 to 0.90 %, respectively (Morga and Kamińska, 2018). In order to compare the elemental

473 composition of graptolite and vitrinite (or telocollinite), data for vitrinite were collected from the literature

474 (Mastalerz and Bustin, 1997; Ward et al., 2005, 2008). O content negatively correlates with C content

475 (Fig. 17a), similar to vitrinite (Mastalerz and Bustin, 1997; Ward et al., 2005, 2008). For elemental

476 comparison with reflectance in Fig. 17, the vitrinite maximum reflectance reported by Ward et al. (2005,

477 2008) was converted to VRor based on the equation proposed by Komorek and Morga (2002), and GRor

478 in Morga and Kamińska (2018) was calculated to EqVRo based on equation (1). C and O content of

479 graptolites is similar to that of vitrinite with similar thermal maturity (Fig. 17b and c) (Mastalerz and

480 Bustin, 1997; Ward et al., 2005, 2008). Similar to the vitrinite, C content of graptolites in shales from

481 Poland increases as organic maturity increases, while O content decreases as organic maturity increases

482 (Fig. 17b and c). The O/C ratio in the graptolites displays a negative correlation with EqVRo (Morga and

483 Kamińska, 2018), comparable to that of vitrinite (Mastalerz and Bustin, 1997; Ward et al., 2005, 2008)

484 (Fig. 17d). These variations of the elemental composition in graptolites are consistent with the maturation

485 process and/or the loss of the O-bearing functional groups from organic macromolecules during thermal

486 evolution (Taylor et al., 1998; Bustin and Guo, 1999). 30 4 (a) Vitrinite (Ward et al., 2008) (b) Vitrinite (Ward et al., 2005) Vitrinite (Mastalerz and Bustin, 1997) Graptolite (Morga and Kamińska, 2018) 3

20 (%)

o 2

O O (%) EqVR 10 1

0 0 60 70 80 90 100 60 70 80 90 100 487 C (%) C (%)

4 4 (c) (d)

3 3

(%) (%) o

o 2 2

EqVR EqVR

1 1

0 0 0 10 20 30 0 0.1 0.2 0.3 0.4 488 O (%) O/C (%)

489 Fig. 17. Comparison of the chemical composition and reflectance of graptolites and vitrinite. The data for the graptolites are from Morga

490 and Kamińska (2018); data on vitrinite from Mastalerz and Bustin (1997) and Ward et al. (2005, 2008). (a) The negative correlation

491 between O and C; (b) the positive correlation between EqVRo and C; (c) the negative correlation between EqVRo and O; (d) the negative

492 correlation between EqVRo and O/C. The legends in b–d are the same as in a.

493

494 5.3.2 Fourier Transform Infrared Spectroscopy (FTIR)

495 FTIR can provide fundamental information not only about the molecular structure of macerals but

496 also about the thermal maturity of the hydrocarbon source rocks, and has been widely used to study the

497 functional groups of the macerals in coals with variable rank (Mastalerz and Bustin, 1993a, b, 1995, 1996;

498 Chen et al., 2012; Wang et al., 2014; Chen et al., 2015b; Wang et al., 2017b). However, several studies

499 have also been conducted on graptolites using FTIR (Bustin et al., 1989; Liu et al., 1996; Suchý et al., 500 2002; Suchý et al., 2004; Caricchi et al., 2016; İnan et al., 2016; Morga and Kamińska, 2018). Bustin et

501 al. (1989) used FTIR to study evolution of graptolites during artificial thermal-treatment, and found that

502 graptolites display a decrease of aliphatic chains and a depletion of aromatic C–H with increase in thermal

503 maturity/heat treatment due to dealkylation and aromatization. Bustin et al. (1989) also found that

504 graptolites are highly aromatic and have fewer aliphatic structures than vitrinite, whereas Hoffknecht

505 (1991) proposed that the percent of aliphatic compounds in graptolites is greater than that in vitrinite.

506 The parameter CH2/CH3 band intensity ratio can provide information about the length of aliphatic

507 chains (Lin and Ritz, 1993), and Suchý et al. (2002) found that the CH3/CH2 ratio increases with

508 increasing GRor. The CH3/CH2 ratio reported by Suchý et al. (2002) were converted to the CH2/CH3 ratio,

509 plotted their data in Fig. 18 and compared it with the results of Morga and Kamińska (2018). The data

510 from these two studies displays opposite correlations of CH2/CH3 ratio with GRor (Fig. 18). The data of

511 the Silurian deposits in the Polish Baltic Basin from Caricchi et al. (2016), however, did not show any

512 obvious correlation (Fig. 18), which may be because in that study, FTIR was conducted on the kerogen

513 concentrate rather than pure graptolites, and the reflectance data was very scattered in a short interval of

514 the same well. In low rank coals (Ro < 1.5%), the CH2/CH3 ratio shows a strong negative correlation with

515 vitrinite reflectance, similar to the results of Morga and Kamińska (2018). This indicates that the aliphatic

516 chains become shorter with maturation. The CHar/CH2+CH3 ratio increases with burial depth and GRor,

517 suggesting an increase in the aromatization of graptolite periderm (Morga and Kamińska, 2018).

518 Graptolites have been thought to be primarily type II kerogen and less commonly type III according to

519 FTIR parameters of ‘A2’ and ‘C2’ factors (Morga and Kamińska, 2018 and references therein).

3

2.5

2 (%)

or 1.5 GR

1

0.5 Suchý et al., 2002 Caricchi et al., 2016 Morga and Kamińska, 2018 0 0.4 0.8 1.2 1.6 2 520 CH2/CH3 521 Fig. 18. The relationship between CH2/CH3 and GRor.

522

523 5.3.3 Raman spectroscopy

524 Raman spectrum is a non-destructive and rapid microstructure analysis technique, and has been used

525 widely as an indicator for thermal maturity of sediments in last several decades (Zerda et al., 1981; Green

526 et al., 1983; Jehlička and Bény, 1992; Wopenka and Pasteris, 1993; Kelemen and Fang, 2001; Romero-

527 Sarmiento et al., 2014; Wilkins et al., 2014; Wilkins et al., 2015; Lünsdorf, 2016; Cheshire et al., 2017;

528 Henry et al., 2018; Jubb et al., 2018; Wilkins et al., 2018; Khatibi et al., 2018a, b). The Raman spectrum

529 of OM is generally composed of two broad peaks, the graphite band (G band) and the disordered band (D

530 band). In the recent decade, this technique has also been used to study graptolites (Suchý et al., 2004; Liu

531 et al., 2013; İnan et al., 2016; Mumm and İnan, 2016; Cheshire et al., 2017; Morga and Pawlyta, 2018;

532 Wang et al., 2019a). Suchý et al. (2004) used micro-Raman spectroscopy to determine the microstructure

533 of the graptolite materials located around an igneous sill, and proposed that the parameter AD/AG

534 (1350/1600 cm–1) peak area can be used to infer organic maturity. Liu et al. (2013) analysed the Raman

535 spectrum from a Silurian graptolite-bearing sediment from Qiaokou, Sichuan, China. Mumm and İnan

536 (2016) and İnan et al. (2016) have conducted Raman spectroscopy on Silurian Qusaiba samples with GRor

537 ranging from 0.76% to 2.20%, and found that both G band position and Raman band separation (RBS;

538 between G peak shift and D peak shift) display strong positive correlations with GRor. Cheshire et al.

539 (2017) have also used RBS to estimate the thermal maturity of the Silurian Qusaiba samples with EqVRo

540 ranging from 0.9% to 2.1% according to the relationship between RBS and VRor proposed by Sauerer et

541 al. (2017), which is consistent with the results from other proxies, (e.g., kerogen elemental composition,

542 kerogen skeletal density and BET specific surface area). Morga and Pawlyta (2018) found that Raman

543 band intensity ratio (ID1/IG) shows a robust correlation with GRor in the Polish Silurian shales, and can be

544 used to calculate EqVRo values. The relationship between ID1/IG and EqVRo has been determined by

545 Morga and Pawlyta (2018) as equation (3):

546 EqVRo = 1.7319 (ID1/IG) – 0.3295 (3)

547 Wang et al. (2019a) studied Raman spectra of graptolites in Chinese sediments from a wider thermal

548 maturity range (GRor ranging from 1.21 to 4.91%). They found that D1 peak position decreases with

549 increasing GRor when GRor < 4.0–4.5%, whereas D1 peak position increases with increasing GRor when

550 GRor > 4.0–4.5%. Both G peak position and RBS display a reverse trend with increasing GRor in 551 comparison with D1 peak position (Wang et al., 2019a). However, it is worth noting that samples with

552 GRor > 4.5% are very limited and the maximum of GRor in this sample set is only 4.91%, thus, more data

553 are needed to illustrate their variation tendency at the high-maturity end. Wang et al. (2019a) established

554 the relationship between GRor and EqVRo as equations (4) and (5), respectively, through the intermediate

555 conversion of RBS (2.1% < GRor < 3.5%) and ID1/IG (GRor > 3.5%) based on the relationship between

556 VRor and RBS (2.1% < GRor < 3.5%), ID1/IG (GRor > 3.5%) proposed by Liu et al. (2013).

557 EqVRo = 0.97 GRor– 0.2 (2.1%< GRor< 3.5%) (4)

558 EqVRo = 0.22 GRor+ 2.55 (GRor> 3.5%) (5)

559 Recently, Hao et al. (2019) compared the Raman spectral parameters of the naturally and artificially

560 matured samples, and the latter has a lower degree of coalification than the former due to insufficient

561 structural transformation of artificially matured samples. They found that GRor displays positive

562 correlations with the full width at half maximum of the D1 band and the G band (FWHM-D/FWHM-G),

563 AD/AG and RBS, and negative correlations with D1 peak position and FWHM-G, and established the

564 relationship between EqVRo and RBS as follows:

565 EqVRo = 0.089 RBS – 19.937 (6)

566 In order to compare the relationship between RBS and reflectance among various types of OM, the

567 data for vitrinite, inertinite, bitumen and graptolite were collected from several publications noted in Fig.

568 19 (Kelemen and Fang, 2001; Suchý et al., 2004; Wilkins et al., 2014; Mumm and İnan, 2016; Sauerer et

569 al., 2017; Morga and Pawlyta, 2018; Wang et al., 2019a; Hao et al., 2019). The RBS of vitrinite, inertinite,

570 bitumen and graptolites increases as thermal maturity increases when the reflectance is lower than 4%,

571 with the exception of the data from Morga and Pawlyta (2018). However, these data become significantly

572 scattered when the reflectance is higher than 4%, especially for the coal/vitrinite data from Liu et al. (2013)

573 and Kelemen and Fang (2001). In addition, G peak position is negatively correlated with GRor in samples

574 from Poland (Morga and Pawlyta, 2018), which opposites to results from the Saudi Arabian Qusaiba

575 samples and the Chinese Wufeng–Longmaxi shales (İnan et al., 2016; Mumm and İnan, 2016; Wang et

576 al., 2019a). These inconsistencies may be due to differences in conditions as follows: (1) sample

577 preparation; (2) experimental conditions (e.g., laser wavelength); (3) data processing; (4) analysis of

578 various OM, and intra-particle chemical heterogeneity (Lünsdorf, 2016; Sauerer et al., 2017; Henry et al.,

579 2018; Jubb et al., 2018; Hao et al., 2019). Thus, uniform conditions are required to better determine the

580 correlation between Raman spectral parameters and graptolite reflectance in future studies. 290

260

)

1

- RBS(cm

230 Coal (Kelemen and Fang, 2001) Vitrinite & inertinite (Wilkins et al., 2014) Vitrinite (Liu et al., 2013) Type III kerogen (Kelemen and Fang, 2001) Type II kerogen (Sauerer et al., 2017) Type II kerogen (Kelemen and Fang, 2001) Graptolite (Suchý et al., 2004) Graptolite (Mumm and İnan, 2016) Graptolite (Morga and Pawlyta, 2018) Graptolite (Wang et al., 2019a) Graptolite (Hao et al., 2019) Bitumen (Liu et al., 2013) 200 0 1 2 3 4 5 6 7 8 581 Ro (%)

582 Fig. 19. The relationship between RBS and reflectance for the vitrinite, inertinite, bitumen and graptolite. The minimum and maximum

583 values of vitrinite and bitumen random reflectance have been given by Liu et al. (2013) rather than the average, thus, half of the total

584 minimum and maximum values were used as Ro in this figure.

585

586 6. The microstructure of graptolites 587 6.1 Scanning Electron Microscopy (SEM) 588 Loucks et al. (2012) classified three types of shale pore systems based on SEM: OM, interparticle,

589 and non-OM intraparticle pores. Although OM pores are within OM, they can form an organic

590 interconnected pore network due to OM connectivity, thus making a significant contribution to the overall

591 porosity (Passey et al., 2010; Loucks et al., 2012; Mastalerz et al., 2013; Milliken et al., 2013). OM pores

592 in the size range from 5 to 750 nm can adsorb and store methane simultaneously (Loucks et al., 2009),

593 and shales with higher OM content display stronger capacity of methane adsorption because of larger

594 surface area supported by OM micropores (Hickey and Henk, 2007; Ross and Bustin, 2007). OM pores

595 are widely identified in North American and Chinese gas-bearing shales, e.g., the Barnett, , Antrim,

596 Marcellus, Woodford, Wufeng–Longmaxi and Longtan shales (Hill and Nelson, 2000; Curtis, 2002; Jarvie

597 et al., 2007; Chalmers et al., 2012; Curtis et al., 2012; Loucks et al., 2012; Mastalerz et al., 2013; Milliken 598 et al., 2013; Tian et al., 2013; Cardott et al., 2015; Hackley and Cardott, 2016; Luo et al., 2016; Ma et al.,

599 2016; Gentzis et al., 2017; Zhang et al., 2017). Pores within graptolites have been studied by SEM (Luo

600 et al., 2016; Ma et al., 2016; Cardott and Curtis, 2018). While many micro-nanopores are widely found

601 in the overmature Wufeng–Longmaxi Formations (Fig. 20) (Luo et al., 2016; Ma et al., 2016), scarce

602 nanopores are present in immature graptolites (%GRor=0.66) from the Lower Ordovician Polk Creek

603 Shale (Cardott and Curtis, 2018). Micro-nano pores in Chinese graptolites generally display irregular or

604 elliptical or nearly spherical shapes, and can be more easily observed in sections parallel to bedding than

605 in sections perpendicular to bedding (Luo et al., 2016; Ma et al., 2016).

606 Graptolites fragments with pores were recently thought to be solid bitumen due to their presence

607 within interparticle pores (İnan et al., 2018). However, Luo et al. (2018) found some graptolites within

608 the interparticle pore space (Fig. 5 in Luo et al., 2018), which look like solid bitumen, but morphologically

609 display graptolitic fusellar layers under polarized light. In order to resolve this conflict, graptolites were

610 marked using diamond lenses during observation under reflected light and then observed under SEM.

611 Many pores were found in the graptolites, although they were smaller and less numerous than those within

612 solid bitumen (Fig. 20).

613 The development and distribution of OM pores in some graptolites are controlled by their fine

614 biological structures (Luo et al., 2016; Ma et al., 2016). Graptolites are generally thought to have low

615 porosity, ranging from 1.62% to 4.23% and averaging 2.51% (Ma et al., 2016). Pores in graptolites can

616 further make up an interconnected system due to their alignment along the fusellar layers, and may

617 connect with discrete porous OM detritus and/or other pore systems (Luo et al., 2016; Ma et al., 2016),

618 which is beneficial to the storage and the exploitation of shale gas in graptolite-bearing sediments. OM

619 pores are also regarded as an important pore system in other gas bearing shales, e.g., (Loucks

620 et al., 2009; Chalmers et al., 2012).

621 622

623 Fig. 20. OM Pores in the Wufeng–Longmaxi Formations. (a) Solid bitumen and graptolite under dry objective; (b) same graptolite under

624 SEM as (a); (c) SEM image showing of location of images d-f; (d-f) images show the pores in solid bitumen are slightly larger and more

625 numerous than that in the graptolite (marked by arrows).

626

627 6.2 High resolution transmission electron microscopy (HRTEM) 628 HRTEM can study OM microstructure at nm scale (Krzesińska et al., 2009; Chalmers et al., 2012;

629 Pawlyta, 2013; Romero-Sarmiento et al., 2014; Morga and Pawlyta, 2018). Morga and Pawlyta (2018)

630 studied the nanostructure of graptolites with GRor of 1.83% using HRTEM. The parallel alignment of

631 carbon layers in the graptolite can be observed in their HRTEM photos, and the dimension of the basic 632 structural units is around 1–2 nm, indicating high molecular ordering, in agreement with the study of

633 Goodarzi (1984).

634 7. The regional maturation and hydrocarbon generation potential of the graptolite-

635 bearing sediments 636 7.1 The hydrocarbon generation potential of graptolites

637 Wang et al. (2017a) concentrated mature (EqVRo around 1.10%) graptolite fragments from two

638 Wufeng–Longmaxi shales. The TOC in the two isolated graptolites is 42.93% and 71.34%, S1 is 6.55 and

639 15.18 mg/g, S2 is 31.71 and 60.36 mg/g, and HI is 74 mg/g TOC and 85 mg/g TOC, respectively, which

640 is much higher than that in the whole rocks (Table 3). İnan et al. (2016) found that an immature isolated

641 graptolite graptolite had HI of 200 mg/g TOC and OI of 30 mg/g TOC, much higher than that of mature

642 graptolites (Wang et al., 2017a). These data indicate that the immature graptolites are hydrogen rich,

643 similar to kerogen type II-III, and have good hydrocarbon potential which is consistent with finding of

644 Hoffknecht (1991), Bustin et al. (1989) and Liu et al. (1996) based on FTIR. Pyrolysis chromatography

645 of the two Chinese graptolies found that the CH4 yields are 12.35 mg/g and 23.61 mg/g, respectively

646 (Wang et al., 2017a). The weak brown fluorescence in the graptolites of the Liteň Formation from the

647 Czech Republic also indicates that the graptolites have some hydrocarbon generation potential (Fig. 5),

648 which means that the graptolite can not be classified as “inertinite-like” macerals (Hoffknecht, 1991).

649 Further, the formation of the OM pores is related with hydrocarbon generation (Loucks et al., 2009),

650 which can explain the rare nanopores in low-maturity graptolites from the Lower Ordovician Polk Creek

651 Shale (Cardott and Curtis, 2018). On the other hand, the abundant pores in the graptolites with high

652 maturity also supports their gas generation and storage potential (Fig. 20). These factors suggest that these

653 graptolites are mainly gas prone, which is consistent with the results of Hoffknecht (1991), and that

654 graptolites possibly make a significant contribution to the contents of shale gas in graptolite-bearing

655 sediments.

656 Table 3. Comparison of the Rock-Eval data of whole rock versus graptolites (MB-2 and MB-3 from Wang et al., 2017a; K1 from İnan et

657 al., 2016).

HI OI Sample ID Lithology Formation S1 (mg/g) S2 (mg/g) Tmax (℃) TOC (%) (mg/g TOC) (mg/g TOC)

MB-2 (whole rock) Shale O3w 0.54 1.77 455 3.04 58 8

MB-3 (whole rock) Shale S1l 0.69 2.88 458 3.88 74 5

MB-2 (graptolite) Graptolite O3w 6.55 31.71 464 42.93 74 43 MB-3 (graptolite) Graptolite S1l 15.18 60.36 456 71.34 85 2

K1 (graptolite) Graptolite S1l 200 30

658

659 Following are examples of worldwide regional maturation and hydrocarbon generation potential

660 studies on graptolite-bearing sedimentary strata. 661 7.2 Canada 662 Graptolite reflectance was used in several regional thermal maturation studies in Canada (Riediger

663 et al., 1989; Link et al., 1990; Goodarzi et al., 1992c; Gentzis et al.,1996). Most of these regional studies

664 were carried out in Canadian Arctic (Goodarzi et al.,1992c; Gentzis et al.,1996) on graptolitic shale and

665 equivalent late Ordovician-early Middle Devonian strata over a distance of over 600 km to determine

666 horizontal variation of thermal maturity and depositional history from Melville Island in southwest to

667 Ellesmere Island in northeast (Fig. 21). The Middle Devonian is up to 3000 m thick, and extends from

668 central Ellesmere Island to Melville Island (Fig. 21) (Trettin, 1989, 1990).

669 Based on graptolite reflectance, the sedimentary succession is overmature (3.5–4.2% GRomax) in

670 western Melville Island, mature in central Melville island (1.75–2.1% GRomax), and immature (0.60–0.80%

671 GRomax) in the Prince of Wales islands, eastern Bathurst, Cornwallis, Baillie Hamilton, Dundas, and

672 Devon Island (Fig. 21). Therefore, the maturity of graptolitic shale increases from mature (1.32% GRomax)

673 in south of Ellesmere Islands to overmature (4.7% GRomax) in north (Fig. 21).

674 Variations in thermal maturity may be due to the areal difference in geothermal gradient and

675 overburden thicknesses. Since regional geothermal gradient has been determined to be ~25 °C/Km in

676 Melville Island (Gentzis, 1991), and this value is considered within conventional scope of geothermal

677 gradient from other North American basins (Gretener, 1982), the difference in thermal maturation is

678 probably due to the thicknesses of the original overburden (Fig. 22). 679

680 Fig 21. Map of the Canadian Arctic Islands showing graptolite maximum reflectance (%GRomax) and maturity of the graptolitic-bearing 681 strata (modified after Gentzis et al., 1996). The green area on the map are designated national park and wilderness. 682

GRomax (%) 0 1 2 3 4 5 0

1000

2000

3000

4000

5000 Erodedsection (m)

6000

7000

683 8000

684 Fig. 22. Variation of graptolite (%GRomax) and overburden strata (eroded sedimentary loading) in Arctic Canada (after Gentzis et al., 685 1996). 686 687 7.3 China 688 The Wufeng–Longmaxi graptolite-bearing sediments in China have various thermal maturities (Luo

689 et al., 2016; Luo et al., 2017; Luo et al., 2018; Wang et al., 2019a), and have the focus of interest due to

690 both the increased importance of shale gas in the global petroleum industry (Curtis, 2002; Bowker, 2007; 691 Jarvie et al., 2007; Ross and Bustin, 2009; Loucks et al., 2012; Dai et al., 2014; Hackley and Cardott,

692 2016), and the successful exploration and exploitation of shale gas in the area (Dai et al., 2014; Dai et al.,

693 2016). The Wufeng–Longmaxi sediments are characterized by abundant OM (TOC: 0.51–25.73%,

694 average 2.59%), overmature levels (mostly EqVRo > 2%), thick beds (30–130 m), abundant OM pores

695 and strong gas generation intensity (Table 4) (Zou et al., 2011; Dai et al., 2014; Dai et al., 2016; Luo et

696 al., 2016; Ma et al., 2016; Luo et al., 2017; Luo et al., 2018), important for the generation and

697 accumulation of shale gas. The first large shale gas field was discovered in the Wufeng–Longmaxi

698 Formations, Fuling, Chongqing, with gas reserves over 100×109 m3 (Dai et al., 2014; Dai et al., 2016).

699 PetroChina developed the Changning-Weiyuan shale gas field in the Wufeng–Longmaxi sediments in

700 2014, and the gas production reached 28×108 m3 in 2016 (Ma and Xie, 2018).

701 In the Wufeng–Longmaxi Formations, graptolite-derived OM accounts for 20–93 vol.% of total OM

702 (Luo et al., 2016). The abundance of graptolites displays a positive correlation with TOC in both Wufeng–

703 Longmaxi Formation from Sichuan Basin and Pingliang Formation from Erdos Basin (Zhu et al., 2015;

704 Deng et al., 2016; Borjigin et al., 2017). Based on the comparison of natural and artificial maturation

705 samples of Wufeng–Longmaxi Formations and Alum Shales (Sweden) (Luo et al., 2018), OM in the high-

706 maturity Wufeng–Longmaxi shales is very similar to the artificially-matured Alum Shales (Fig. 23). This

707 suggests that OM in immature Wufeng–Longmaxi shales is similar to OM in the immature Alum Shales.

708 The Wufeng–Longmaxi Formations have been subdivided and correlated on the basis of lithology

709 or logging curves, however, it has not been accepted widely due to the lower resolution and precision of

710 these methods (Zou et al., 2015; Chen et al., 2017). As important biostratigraphic indicator, graptolite

711 species have been widely used for the stratigraphic correlation (Chen, 1984; Bergström and Mitchell,

712 1986; Chen et al., 1987; Cooper, 1999; Chen et al., 2000; Chen et al., 2004; Chen et al., 2005; Loydell,

713 2011; Chen et al., 2017). Chen et al. (2015a) proposed a new subdivision and correlation for the Wufeng–

714 Longmaxi Formations based on the graptolite biostratigraphy. According to this revised stratigraphy,

715 intervals (WF2–WF3 and LM2–LM6) were regarded as the best targets for shale gas (Chen et al., 2015a;

716 Chen et al., 2017), and they generally contain abundant graptolites (> 30%) (Qiu et al., 2018). The pores

717 in these graptolites also act as a reservoir and/or migration pathway of the shale gas (Ma et al., 2016; Luo

718 et al., 2018; Qiu et al., 2018).

719 720 721 Fig. 23. Comparison of OM in natural Wufeng–Longmaxi shales (a) and heat-treated Alum Shale heated at 550 ℃ for 3 days (b).

722

723 A regional maturation pattern based on GRor converted to vitrinite (Yang, 2016; Luo et al., 2018;

724 Wang et al., 2019a) for the Wufeng–Longmaxi Formations from Chongqing and surrounding areas

725 indicates that graptolite-bearing strata are mostly overmature, with maturity increasing from southwest

726 towards northeast with an EqVRo of 1.17–4.93%, the maximum in sample TT from Pingshan, Sichuan

727 province, and minimum in well YC1 from Chengkou, Chongqing (Fig. 24). 728 Table 4. Geological and geochemical data of selected global graptolite-bearing sediments.

Country/Continent Area/Basin Formation Age TOC (%) EqVRo (%) Depth (m) Thickness (m) Porosity (%) Gas content (m3/t)

China Chongqing Wufeng–Longmaxi Upper Ordovician–Lower Silurian 0.51–25.73 1.17–4.93 900–4500 30–130 3–10 1.7–4.5 Norwegian– Denmark Danish Basin Alum Shale Miaolingian– 3.00–9.00 1.8–2.5 2000–4000 30–180 3–6 avg. 0.85 (onshore) Sweden Colonus Trough Alum Shale Miaolingian–Tremadocian 3.00–16.00 1.7–2.0 700–1000 70–90 avg. 6.5% avg. 0.85

Swedena Central Sweden Alum Shale Miaolingian–Tremadocian 1.00–21.00 0.5–1.1 20–100 20–35 / max. 7.6

Poland BPLBb Piaśnica Upper –Tremadocian 3–12 0.55–2.40c /d 10–34 3.17–6.89 / Poland BPLB Sasino Caradoc 1–7 0.5–4.2c / 1.5–70 2.24–10.94 / Poland BPLB Pasłęk Llandovery 1–6 0.7–4.0c / 20–80 3.95–5.56 (Jantar Member) / Poland BPLB Pelplin–lower part Wenlock 0.5–1.7 0.6–2.5c / 30–1000 5.25–7.88 / North Africa Ghadames Basin Tanezzuft Llandovery–Wenlock 2–16 0.7–2.0 2000–5000 20–100 / 2.4–8.5

729 a Biogenic gas;

730 b BPLB = Baltic–Podlasie–Lublin Basin;

c 731 Mean random reflectance of the vitrinite–like macerals (Rvl);

732 d “/” means no data.

733 734

735 Fig. 24. The EqVRo distribution of the Wufeng–Longmaxi shales in Chongqing and surrounding regions. Data at each sample location is

736 well/outcrop name/EqVRo. Data are from Yang (2016); Luo et al. (2018); Wang et al. (2019a).

737

738 7.4 Denmark and Sweden 739 Lower Palaeozoic strata is widely distributed in Denmark, Sweden, Norway, Poland and the Baltic

740 States (Fig. 25). In this area, terrestrial to shallow marine sandy deposition commenced in lower Cambrian

741 (Serie 2) following an overall transgression of the Baltic continent (Nielsen and Schovsbo, 2011, 2015).

742 From the middle Cambrian (Miaolingian) to the Lower Ordovician (Tremadocian), organic rich mud

743 deposition known as the in Denmark, Sweden and Norway, the Türisalu Formation

744 in Estonia, the Kaporye Formation in the St. Petersburg region and the Slowinski Formation as well as

745 Piaśnica Formation in NE Poland completely dominated (Nielsen and Schovsbo, 2007). The shale is now

746 present in the deeply buried strata towards the Polish-German and Swedish-Norwegian Caledonian fronts, 747 in the Baltic area and as erosional outliers in Sweden (Fig. 25), representing only a minor part of the

748 original huge deposition area that its maximum at the Lower Ordovician may have extended for more

749 than 800 000 km2 (Schovsbo et al., 2018).

750

751

752 Fig. 25. Outline of approximate original distribution of the Alum Shale Formation and present-day occurrence of the Lower Paleozoic

753 strata in southern Scandinavia (modified from Schovsbo et al., 2018).

754

755 Thermogenic gas trapped in the Alum Shales has been explored in Denmark (Schovsbo and Jakobsen,

756 2019) and Sweden (Pool et al., 2012) (Table 4). In Denmark and Sweden, a considerable resource has

757 been estimated by the U.S. Geological Survey (mean = 67 × 109 m3 gas; Gautier et al., 2013) and by the

758 European Geological Surveys (Zijp et al., 2017) to be present mostly in the Cambrian (Miaolingian) to

759 Ordovician (Tremadocian) Alum Shale Formation. Typical Alum Shale has TOC values from 5–10%

760 (even up to 25% TOC in some samples) (Fig. 26), and their thermal maturity ranges from immature

761 (EqVRo<0.5%) in Central Sweden and Estonia, dry gas mature (EqVRo: 1.6–2.5%) in most of Denmark

762 and southern Sweden, and to post-mature to low-grade metamorphic in Norway (EqVRo >3%; Buchardt

763 et al., 1997; Petersen et al., 2013) (Table 4). The shale is graptolitic with macroscopic remains of typical

764 the pelagic Rhapdinopora occurring abundantly in the Tremadocian shale. In the Cambrian, microscopic

765 remains of graptolites also occur presumably from benthic species (Petersen et al., 2013) - an 766 interpretation that has been supported by the rediscovery of Miaolingian benthic graptolites in the Alum

767 Shale from Norway (Wolvers and Maletz, 2016). At this time, no quantification of the contribution of

768 graptolite carbon to the total TOC content have been made, and the only published study of the Alum

769 Shale nanopore system by Henningsen et al., 2018 did not specifically investigate the graptolite

770 contribution to the total porosity.

771 The Lower Paleozoic shale gas prospective areas in Denmark largely follow the margins of the

772 Norwegian–Danish Basin (Schovsbo et al., 2014). Exploration of this play is still limited, and

773 representative well data and production test data for the shales are lacking. The only exploration well in

774 Denmark revealed gas in the Alum Shale Formation, however, no test production was carried out, so the

775 commercial potential is still unknown (Schovsbo and Jacobsen, 2019). The well was drilled within a so-

776 called ‘sweet spot’, defined as an area with expected highest gas content (c.f. Schovsbo et al., 2014).

777 Results from this well revealed the Alum Shale to be 40 m thick, compared to up to 180 m in the

778 depocenter offshore Denmark. In Sweden, thermogenic gas has been explored in Scania, within the

779 Colonus Trough, which is a fault bounded graben system in southern Sweden. Exploration indicated that

780 the Alum Shale Formation was gas mature (EqVRo: 1.7–2%), located at 700–1000 m depth (Table 4), but

781 did not contain gas in economically producible quantities, possibly due to gas leakage as a result of uplift

782 (Pool et al., 2012). In south-central Sweden, gas-bearing immature to low-maturity Alum Shale with

783 depths < 150 m have been known for several decades and been under exploration for commercial shale

784 gas production (Table 4); they are thought to be composed of mixtures of thermogenic and bacterially-

785 derived gas (Schultz et al., 2015; Schovsbo and Nielsen 2017). 786

787 Fig. 26. Composite profile of TOC content in the Alum Shale in Scania, southern Sweden (modified from Schovsbo, 2003).

788 7.5 Poland 789 The highest shale gas exploration potential in Poland is related to the Cambrian–Lower Ordovician

790 (upper Furongian–Tremadocian), Upper Ordovician (Caradoc) and Silurian (Llandovery–Wenlock) black

791 and dark grey shales with abundant graptolite remains within the Baltic-Podlasie-Lublin Basin (BPLB) at

792 the slope of the East European Craton (EEC) (Table 4, Figs. 27 and 28) (Poprawa, 2010; Karcz et al.,

793 2013; Podhalańska, 2013). Due to lateral facies variability, hydrocarbon potential of these rocks varies

794 throughout the area. The Lower Paleozoic deposits were buried to significantly different depths in the

795 individual parts of the BPLB. Tectonic processes resulted in a break-up of the Basin into segments (Karcz

796 et al., 2013). The current burial depth within the significant part of BPLB is shallow enough for

797 commercial shale gas exploration, and geological structure is simple, especially within the Baltic Basin

798 and Podlasie Basin (Poprawa, 2010). 799

800 Fig. 27 The distribution of Upper Ordovician–Lower Silurian sediments in the Baltic-Podlasie-Lublin Basin (BPLB) at the slope of the

801 East European Craton (EEC) (modified after Poprawa, 2010). Abbreviations: SPW =Płock-Warsaw zone; SBN = Biłgoraj-Narol zone;

802 TESZ = Trans-European Suture zone.

803

804 The upper Furongian–Tremadocian Alum-Shale-equivalent black bituminous shales belong to the

805 Piaśnica Formation, which occurs only within the Baltic Basin (Fig. 28). Their thickness is from 10 m (in

806 the onshore part) to 34 m (in the offshore part) (Szymański, 2008; Więcław et al., 2010; Podhalańska et

807 al., 2016a), thinner than the Alum Shale in Denmark (30–180 m) and southern Sweden (70–90 m); and

808 average TOC content ranges between 3% and 12.0% (Table 4). Thermal maturity of OM increases with

809 the increasing depth from NE to SW from the main phase of oil generation through the condensate and

810 wet gas phase to the dry gas generation phase. The Rock-Eval Tmax temperature and Rvl vary from 430 °C 811 to 505 °C, and from 0.55% to 2.4%, respectively (Poprawa, 2010; Więcław et al., 2010; Karcz and Janas,

812 2016; Podhalańska et al., 2016a) (Table 4). The effective porosity ranges between 3.17% and 6.89%

813 (Table 4) (Dyrka, 2016). S2 is up to 72 mg HC/g rock, while HI varies from 6 to 484 mg HC/ TOC

814 (Więcław et al., 2010). The gas content in the Piaśnica Formation reaches 7.6 m3/t in the Lębork S-1 well

815 (Lehr and Keeley, 2016) (Table 4), comparable to that in North American gas-bearing sediments (e.g.

816 Jarvie, 2012). 817

818 Fig. 28 Generalized lithostratigraphy of the Upper Cambrian-Silurian deposits in the Polish part of the Baltic-Podlasie-Lublin Basin

819 (after Podhalańska 2016b). 820 The Caradoc (Sandbian–lower Katian) rocks (i.e., Sasino Formation) are most fully developed in the

821 northern and western parts of the Baltic Basin, mainly composed of dark grey and black shales with rich

822 graptolite fauna (Fig. 28) (Modliński and Szymański, 1997; Podhalańska, 2013; Podhalańska et al.,

823 2016a). Similar rocks but with lower thickness and stratigraphic extent stretch to the southeast in the

824 Podlasie and Lublin areas (different regional lithostratigraphic equivalents). Thickness of the Sasino

825 Formation and its equivalents in other parts of the BPLB varies from 1.5 m in the eastern onshore part of

826 the Basin to 70 m in its offshore part (Table 4) (Modliński and Szymański, 1997, 2008; Poprawa, 2010;

827 Więcław et al., 2010; Podhalańska et al., 2016a; Podhalańska et al., 2016b). The average TOC content in

828 Sasino Formation sediments ranges from 1% to 4%, and even up to 7%, in the offshore part (Table 4)

829 (Poprawa, 2010; Więcław et al., 2010; Karcz and Janas 2016; Podhalańska et al., 2016a). Thermal

830 maturity increases from NE to SW with increasing depth from the main phase of oil generation (0.6–1.1%

831 Rvl) through condensate and wet gas generation phase (1.1–1.4% Rvl) to the dry gas generation phase

832 (>1.4% Rvl), reaching 4.2% within the Baltic Basin (Table 4) (Więcław et al., 2010; Podhalańska et al.,

o 833 2016a). Tmax values reported for the BPLB are 425–474 C (Więcław et al., 2010; Karcz and Janas 2016).

834 The effective porosity ranges between 2.24% and 10.94% (Table 4) (Dyrka, 2016). S2 reported for the

835 Baltic Basin varies from 0.05 to 10.6 mg HC/g rock, and HI ranges from 11 to 359 HC/g TOC.

836 Llandovery (Rhuddanian-) shales make up the Pasłęk Formation, whereas the Wenlock

837 (-) rocks belong to the lower part of the Pelplin Formation (Sheinwoodian-

838 as a whole) (Fig. 28) (Podhalańska et al., 2016b). They contain rich and variable graptolite

839 fauna, which made it possible to determine biostratigraphic levels and limits of chronostratigraphic units

840 (Podhalańska, 2013). The Llandovery shales of the Pasłęk Formation occur throughout vast parts of the

841 western slope of EEC. Their thickness generally increases from the east to the west, reaching a maximum

842 of 80 m. However, in the major part of the BPLB, it ranges from 20 to 40 m (Table 4). Average TOC

843 values are usually 1% to 2.5%, except for the Podlasie Basin, where they reach 6% (and even > 15% in

844 individual layers) (Table 4). High TOC content was particularly documented in the lower part of the

845 Pasłęk Formation (the Jantar member) (Modliński et al., 2006; Poprawa 2010; Więcław et al., 2010; Karcz

846 et al., 2013). Rvl in the Baltic Basin varies from 0.7% to 2.3%, and rises to 4.0% in the Podlasie Basin,

847 showing the NE-SW trend (Table 4) (Poprawa, 2010; Więcław et al., 2010; Karcz et al., 2013). Tmax

o 848 temperature reported for the Baltic Basin falls within the range of 435–534 C. S2 varies from 0.12 to 40.6

849 mg HC/g rock, while HI varies from 8 to 436 mg HC/g TOC (Więcław et al., 2010). 850 The Jantar member (Rhuddanian–lower ), composed of bituminous shales, is the most

851 perspective unit of the Pasłęk Formation within the onshore and offshore areas of the Baltic Basin (Fig.

852 28) (Podhalańska et al., 2016a). Its maximum thickness is 18 m (Modliński et al., 2006; Podhalańska et

853 al., 2016a). The average TOC content varies from 2% to 5%, and the thermal maturity increases from NE

854 to SW, starting with the oil generation window and passing through condensate and wet gas phase to dry

855 gas generation phase (Poprawa 2010; Karcz and Janas, 2016; Podhalańska et al., 2016a; Podhalańska et

856 al., 2016b). Rock-Eval Tmax values range between 438 and 485 °C (Karcz and Janas, 2016). Average

857 effective porosity is 3.95–5.56% (Table 4) (Dyrka, 2016).

858 The Wenlock shales have the largest extent of all stratigraphic units. They have been documented

859 within the whole BPLB (Fig. 28) (Podhalańska et al., 2016a). Their thickness is highly variable, from 30

860 m in SE part of the Lublin Basin to over 1000 m in western part of the Baltic Basin (Table 4) (Poprawa,

861 2010; PIG, 2012; Karcz et al., 2013; Podhalańska et al., 2016a). Average content of OM in individual

862 Wenlock sections in central and western parts of the Baltic Basin and the Podlasie Basin usually ranges

863 from 0.5% to 1.4% TOC. In eastern part of the Baltic Basin and in the Lublin Basin, it is higher, rising to

864 about 1–1.7% TOC (Poprawa, 2010, Więcław et al., 2010; Karcz and Janas, 2016; Podhalańska et al.,

865 2016a). Thermal maturity of OM increases from NE to SW, from the oil generation phase through the

866 condensate and wet gas phases to dry gas generation phase (Podhalańska et al., 2016a). Rvl equals 0.6–

o o 867 2.5% (Table 4), and Tmax values vary from 425 C to 509 C (Więcław et al., 2010; Karcz and Janas 2016;

868 Podhalańska et al., 2016a). Average effective porosity is 5.25–7.88% (Table 4) (Dyrka, 2016). S2 in the

869 Baltic Basin ranges between 0.07 and 3.5 mg HC/g rock. HI is 7–335 mg HC/g TOC (Więcław et al.,

870 2010). 871 7.6 Arabia and North Africa 872 The Lower Silurian organic-rich shales (hot shales) were widely deposited in the Arabia and North

873 Africa due to a strong rise of sea level during the latest Ordovician to Early Silurian, including e.g.,

874 Tanezzuft Formation in Libya, Tunisia and Algeria, Tanf Formation in Syria, Akkas Formation in Iraq,

875 Qusaiba Formation in Saudi Arabia, and Batra or Mudawwara Formation in Jordan (Loydell, 1998;

876 Lüning et al., 2003; Soua, 2014; İnan et al., 2016). They are characterized by type II kerogen, TOC up to

877 20%, and thickness is generally less than 100 m (Lüning et al., 2000; Belaid et al., 2010; Soua, 2014;

878 Wang et al., 2019b). Their thermal maturity varies significantly due to variable burial history, ranging

879 from immature to post mature (Lüning et al., 2005; Belaid et al., 2010; Soua, 2014; İnan et al., 2016; 880 Wang et al., 2019b). In North Africa, 80–90% of Paleozoic-sourced petroleum are derived from these

881 Silurian hot shales, and in Arabia, some oil and sweet gas in Paleozoic-Mesozoic reservoirs are regarded

882 to be sourced from these organic-rich sediments (Lüning et al., 2000; Al-Juboury and Al-Hadidy, 2009;

883 İnan et al., 2016). In recent years, the hot shales were thought to contain shale oil/gas potential due to

884 their wide distribution, thickness, abundant OM and variable thermal maturity (Soua, 2014; İnan et al.,

885 2016; Wang et al., 2019b). For example, the gas content of the hot shales falls between 2.4 and 8.5 m3/t

886 in the Ghadames Basin, and their geochemical characteristics and petroleum accumulation conditions are

887 comparable to the Wufeng–Longmaxi Formations (Table 4) (Wang et al., 2019b). 888 8. Conclusions 889 Based on a review on the optical characteristics, chemical composition, and microstructure of the

890 graptolites, the following conclusions can be reached:

891 (1) There are two types of graptolites: granular (GG) and non-granular (NGG). GG have lower

892 reflectance and weaker anisotropy than NGG. GG can be altered to NGG with increasing thermal maturity.

893 (2) The aromaticity and ordering of the aromatic structure of graptolite increasing with increasing

894 maturity is similar to that of vitrinite and bitumen, illustrating that their physical and chemical parameters

895 are reliable geo-thermometers. There is a positive correlation between GRor and GRomax, allowing GRor

896 to be used as a measure of thermal maturation for the Lower Paleozoic sediments. The relationship

897 between GRor and EqVRo can be expressed as: EqVRo = 0.99GRor + 0.08.

898 (3) Electron microprobe and FTIR analyses are very useful tools to study chemical compositions of

899 graptolites, indicating that their compositions are similar to vitrinite. Standard measurement and data

900 processing are required for the improved analysis of Raman spectrum for graptolites. Graptolites contain

901 abundant pores, which are less abundant and smaller than those in solid bitumen.

902 (4) Based on the chemical composition and organic geochemical characteristics of the graptolites,

903 they are mainly gas prone, similar to vitrinite, and have a significant hydrocarbon generation potential.

904 (5) OM in the Wufeng–Longmaxi sediments is dominated by graptolites and solid bitumen. Their

905 thermal maturity was determined by GRor, in the range of 1.17–4.93%, indicating marginally mature to

906 overmature.

907 (6) The graptolite-bearing shales are worldwide hydrocarbon source rocks and contribute

908 significantly to global petroleum reserves. Graptolite serve both as source material and, due to nano-

909 microporosity, as reservoir and are important to the accumulation of shale gas in these sediments. 910 911 Acknowledgments 912 This work was supported by National Natural Science Foundation of China (No. 41503028,

913 41773031 and 41830424). Field sampling in the Czech Republic was partly supported by OP RDE, MEYS,

914 under the project “Ultra-trace isotope research in social and environmental studies using accelerator mass

915 spectrometry” (Reg. No. CZ.02.1.01/0.0/0.0/16_019/0000728). We thank Dr. Roger MacQueen of the

916 Geological Survey of Canada for reviewing an early version of this manuscript and for his valuable

917 suggestion. We also thank Zhongliang Ma at the Wuxi Institute of Petroleum Geology, Sinopec Petroleum

918 Exploration and Development Research Institute for conducting the artificial maturation experiments in

919 this study. We are very grateful to Dr. MaryAnn Love Malinconico and one anonymous reviewer for their

920 critical but constructive and valuable comments that significantly improved the quality and language of

921 this paper. The authors also appreciate Managing Editor Dr. Shuhab Khan for his precious time and energy

922 to handle this paper.

923

924 References

925 Al-Juboury, A.I., Al-Hadidy, A.H., 2009. Petrology and depositional evolution of the Paleozoic rocks of Iraq. Marine and Petroleum 926 Geology 26, 208–231. 927 Bates, D., Kirk, N., 1986. Mode of secretion of graptolite periderm, in normal and retiolite graptolites. Geological Society Special 928 Publication 20, 221–236. 929 Belaid, A., Krooss, B.M., Littke, R., 2010. Thermal history and source rock characterization of a Paleozoic section in the Awbari Trough, 930 Murzuq Basin, SW Libya. Marine and Petroleum Geology 27, 612–632. 931 Bergström, S., Mitchell, C., 1986. The graptolite correlation of the North American Upper Ordovician Standard. Lethaia 19, 247–266. 932 Bertrand, R., 1990. Correlations among the reflectances of vitrinite, chitinozoans, graptolites and scolecodonts. Organic Geochemistry 15, 933 565–574. 934 Bertrand, R., 1993. Standardization of solid bitumen reflectance to vitrinite in some Paleozoic sequences of Canada. Energy Sources 15, 935 269–287. 936 Bertrand, R., Heroux, Y., 1987. Chitinozoan, graptolite, and scolecodont reflectance as an alternative to vitrinite and pyrobitumen 937 reflectance in Ordovician and Silurian strata, , , Canada. AAPG Bulletin 71, 951–957. 938 Bertrand, R., Lavoie, D., Fowler, M., 2003. Cambrian-Ordovician shales in the Humber zone: thermal maturation and source rock potential. 939 Bulletin of Canadian Petroleum Geology 51, 213–233. 940 Bertrand, R., Malo, M., 2001. Source rock analysis, thermal maturation and hydrocarbon generation in Siluro-Devonian rocks of the Gaspé 941 Belt basin, Canada. Bulletin of Canadian Petroleum Geology 49, 238–261. 942 Borjigin, T., Shen, B., Yu, L., Yang, Y., Zhang, W., Tao, C., Xi, B., Zhang, Q., Bao, F., Qin, J., 2017. Mechanisms of shale gas generation 943 and accumulation in the Ordovician Wufeng–Longmaxi Formation, Sichuan Basin, SW China. Petroleum Exploration and 944 Development 44, 69–78. 945 Bowker, K.A., 2007. Barnett Shale gas production, Fort Worth Basin: issues and discussion. AAPG bulletin 91, 523–533. 946 Briggs, D.E.G., Kear, A.J., Baas, M., Deleeuw, J.W., Rigby, S., 1995. Decay and composition of the Hemichordate Rhabdopleura - 947 implications for the taphonomy of graptolites. Lethaia 28(1), 15–23. 948 Brown, H., Taylor, G., 1961. Some remarkable Antarctic coals. Fuel 40, 211–244. 949 Buchardt, B., Lewan, M.D., 1990. Reflectance of vitrinite-like macerals as a thermal maturity index for Cambrian-Ordovician Alum shale, 950 southern Scandinavia. AAPG Bulletin 74, 394–406. 951 Buchardt, B., Nielsen, A.T., Schovsbo, N.H., 1997. Alun Skiferen i Skandinavien. Geologisk Tidsskrift 3, 1–30 [in Danish]. 952 Bustin, R., Link, C., Goodarzi, F., 1989. Optical properties and chemistry of graptolite periderm following laboratory simulated maturation. 953 Organic Geochemistry 14, 355–364. 954 Bustin, R.M., Guo, Y., 1999. Abrupt changes (jumps) in reflectance values and chemical compositions of artificial charcoals and inertinite 955 in coals. International Journal of Coal Geology 38, 237–260. 956 Bustin, R.M., Mastalerz, M., Wilks, K.R., 1993. Direct determination of carbon, oxygen and nitrogen content in coal using the electron 957 microprobe. Fuel 72, 181–185. 958 Cardott, B.J., Kidwai, M.A., 1991. Graptolite reflectance as a potential thermal-maturation indicator. In Johnson, K.S., (ed.), Late 959 Cambrian-Ordovician geology of the southern Midcontinent, 1989 symposium. Oklahoma Geological Survey Circular 92, 203-209. 960 Cardott, B.J., Curtis, M.E., 2018. Identification and nanoporosity of macerals in coal by scanning electron microscopy. International 961 Journal of Coal Geology 190, 205–217. 962 Cardott, B.J., Landis, C.R., Curtis, M.E., 2015. Post-oil solid bitumen network in the Woodford Shale, USA—A potential primary 963 migration pathway. International Journal of Coal Geology 139, 106–113. 964 Caricchi, C., Corrado, S., Di Paolo, L., Aldega, L., Grigo, D., 2016. Thermal maturity of Silurian deposits in the Baltic Syneclise (on-shore 965 Polish Baltic Basin): contribution to unconventional resources assessment. Italian Journal of Geosciences 135, 383–393. 966 Cartz, L., Hirsch, P.B., 1960. A contribution to the structure of coals from X-ray diffraction studies. Philosophical Transactions of the 967 Royal Society of London. Series A, Mathematical and Physical Sciences 252, 557–602. 968 Chalmers, G.R., Bustin, R.M., Power, I.M., 2012. Characterization of gas shale pore systems by porosimetry, pycnometry, surface area, 969 and field emission scanning electron microscopy/transmission electron microscopy image analyses: examples from the Barnett, 970 Woodford, Haynesville, Marcellus, and Doig units. AAPG Bulletin 96, 1099–1119. 971 Chandra, D., 1963. Reflectance of thermally metamorphosed coals. Fuel 42, 69–74. 972 Chen, X., 1984. Influence of the Late Ordovician glaciation on basin configuration of the Yangtze Platform in China. Lethaia 17, 51–59. 973 Chen, X., Fan, J., Wang, W., Wang, H., Nie, H., Shi, X., Wen, Z., Chen, D., Li, W., 2017. Stage-progressive distribution pattern of the 974 Lungmachi black graptolitic shales from Guizhou to Chongqing, Central China. Science China Earth Sciences 60, 1133–1146. 975 Chen, X., Fan, J., Zhang, Y., Wang, H., Chen, Q., Wang, W., Liang, F., Guo, W., Zhao, Q., Nie, H., Wen, Z., Sun, Z., 2015a. Subdivision 976 and delineation of the Wufeng and Lungmachi black shales in the subsurface areas of the Yangtze platform. Journal of Stratigraphy 977 39, 351–358 (in Chinese with English abstract). 978 Chen, X., Rong, J., Charles, M., David, H., Fan, J., Zhan, R., Zhang, Y., Li, R., Wang, Y., 2000. Late Ordovician to earliest Silurian 979 graptolite and brachiopod biozonation from the Yangtze region, South China, with a global correlation. Geological Magazine 137, 980 623–650. 981 Chen, X., Rong, J., Li, Y., Boucot, A., 2004. Facies patterns and geography of the Yangtze region, South China, through the Ordovician 982 and Silurian transition. Palaeogeography, Palaeoclimatology, Palaeoecology 204, 353–372. 983 Chen, X., Rong, J., Michael, M., David, S., Charles, M., Fan, J., 2005. Patterns and processes of latest Ordovician graptolite extinction 984 and recovery based on data from South China. Journal of Paleontology 79, 824–861. 985 Chen, X., Xiao, C., Chen, H.y., 1987. Wufengian (Ashgillian) graptolite faunal differentiation and anoxic environment in South China. 986 Acta Palaeontologica Sinica 26, 326–344 (in Chinese with English abstract). 987 Chen, Y., Furmann, A., Mastalerz, M., Schimmelmann, A., 2014. Quantitative analysis of shales by KBr-FTIR and micro-FTIR. Fuel 116, 988 538–549. 989 Chen, Y., Mastalerz, M., Schimmelmann, A., 2012. Characterization of chemical functional groups in macerals across different coal ranks 990 via micro-FTIR spectroscopy. International Journal of Coal Geology 104, 22–33. 991 Chen, Y., Zou, C., Mastalerz, M., Hu, S., Gasaway, C., Tao, X., 2015b. Applications of Micro-Fourier Transform Infrared Spectroscopy 992 (FTIR) in the Geological Sciences—A Review. International Journal of Molecular Sciences 16, 26227. 993 Cheshire, S., Craddock, P.R., Xu, G., Sauerer, B., Pomerantz, A.E., McCormick, D., Abdallah, W., 2017. Assessing thermal maturity 994 beyond the reaches of vitrinite reflectance and Rock-Eval pyrolysis: A case study from the Silurian Qusaiba formation. International 995 Journal of Coal Geology 180, 29–45. 996 Clarkson, E.N.K., 1981. Invertebrate Paleontology and Evolution. George Allen & Unwin, Boston. 997 Clausen, C., Teichmüller, M., 1982. Die Bedeutung der Graptolithenfragmente im Paläozoikum von Soest-Erwitte für Stratigraphie und 998 Inkohlung. Fortschritte in der Geologie von Rheinland und Westfalen 30, 145–167. 999 Cole, G.A., 1994. Graptolite-chitinozoan reflectance and its relationship to other geochemical maturity indicators in the Silurian Qusaiba 1000 Shale, Saudi Arabia. Energy & fuels 8, 1443–1459. 1001 Colţoi, O., Nicolas, G., Safa, P., 2016. The assessment of the hydrocarbon potential and maturity of Silurian intervals from eastern part of 1002 Moesian Platform – Romanian sector. Marine and Petroleum Geology 77, 653–667. 1003 Cook, A., Murchison, D., Scott, E., 1972. Optically biaxial anthracitic vitrinites. Fuel 51, 180–184. 1004 Cooper, R.A., 1999. Ecostratigraphy, zonation and global correlation of earliest Ordovician planktic graptolites. Lethaia 32, 1–14. 1005 Crowther, P.R., 1981. The fine structure of graptolite periderm. Special papers in palaeontology, no. 26. Palaeontological Association, 1006 London. 1007 Curtis, J.B., 2002. Fractured shale-gas systems. AAPG Bulletin 86, 1921–1938. 1008 Curtis, M.E., Cardott, B.J., Sondergeld, C.H., Rai, C.S., 2012. Development of organic porosity in the Woodford Shale with increasing 1009 thermal maturity. International Journal of Coal Geology 103, 26–31. 1010 Dai, J., Zou, C., Liao, S., Dong, D., Ni, Y., Huang, J., Wu, W., Gong, D., Huang, S., Hu, G., 2014. Geochemistry of the extremely high 1011 thermal maturity Longmaxi shale gas, southern Sichuan Basin. Organic Geochemistry 74, 3–12. 1012 Dai, J., Zou, C., Dong, D., Ni, Y., Wu, W., Gong, D., Wang, Y., Huang, S., Huang, J., Fang, C., Liu, D., 2016. Geochemical characteristics 1013 of marine and terrestrial . Marine and Petroleum Geology 76, 444–463. 1014 Davis, A., 1978. The reflectance of coal. In: Karr, C. (Ed.), Analytical methods for coal and coal products. Academic Press, London, pp. 1015 27–81. 1016 Deng, K., Zhou, W., Zhou, L., Wan, Y., Deng, H., Xie, R., Chen, W., 2016. Influencing factors of micropores in the graptolite shale of 1017 Ordovician Pingliang Formation in Ordos Basin, NW China. Petroleum Exploration and Development 43, 416–424. 1018 Dyrka, I., 2016. Charakterystyka petrofizyczno–mineralogiczna perspektywicznych kompleksów węglowodorowych na podstawie 1019 wybranych wyników badań laboratoryjnych. Przegląd Geologiczny 64, 982–986 (in Polish with English abstract). 1020 Eisenack, A., 1932. Neue Mikrofossilien des baltischen Silurs. II. Palaeontologische Zeitschrift 14, 257–277. 1021 Epstein, A., Epstein, J., Harris, L., 1977. Conodont colour alteration-an index of organic metamorphism. US Geol. Surv. Prof. Pap. 995, 1022 27. 1023 Ferreiro Mählmann, R., Le Bayon, R., 2016. Vitrinite and vitrinite like solid bitumen reflectance in thermal maturity studies: Correlations 1024 from diagenesis to incipient metamorphism in different geodynamic settings. International Journal of Coal Geology 157, 52–73. 1025 Fink, R., Virgo, S., Arndt, M., Visser, W., Littke, R., Urai, J.L., 2016. Solid bitumen in calcite veins from the Natih Formation in the Oman 1026 Mountains: Multiple phases of petroleum migration in a changing stress field. International Journal of Coal Geology 157, 39–51. 1027 Gautier, D.L., Charpentier R.R., Gaswirth, S.B., Klett, T.R., Pitman, J.K., Schenk, C.J., Tennyson, M.E., Whidden, K.J., 2013. 1028 Undiscovered gas resources in the Alum Shale, Denmark. U.S. Geological Survey Fact Sheet 2013–3103, 4 pp. 1029 Gentzis, T., 1991. Regional maturity and source rock potential of Paleozoic and Mesozoic strata, Melville Island, Arctic Canada. Ph.D. 1030 thesis, University of Newcastle-Upon Tyne, United Kingdom. 1–444. 1031 Gentzis, T., Carvajal-Ortiz, H., Ocubalidet, S., and Wawak, B., 2017. Organic Petrology Characteristics of Selected Shale Oil and Shale 1032 Gas Reservoirs in the USA: Examples from “The Magnificent Nine”. In: The Role of Organic Petrology in the Exploration of 1033 Conventional and Unconventional Hydrocarbons Systems, Suárez-Ruiz, I., Graciano Mendoça Filho, J., (Eds.). Bentham Science 1034 Publishers, Sharjah, UAE, 1, 131-167. 1035 Gentzis, T., de Freitas, T., Goodarzi, F., Melchin, M., Lenz, A., 1996. Thermal maturity of Lower Paleozoic sedimentary successions in 1036 Arctic Canada. AAPG bulletin 80, 1065–1083. 1037 George, S., Ahmed, M., 2002. Use of aromatic compound distributions to evaluate organic maturity of the middle Velkerri 1038 Formation, McArthur Basin, Australia. The Sedimentary Basins of Western Australia 3, 253–270. 1039 George, S., Llorca, S., Hamilton, P., 1994. An integrated analytical approach for determining the origin of solid bitumens in the McArthur 1040 Basin, northern Australia. Organic Geochemistry 21, 235–248. 1041 Gonçalves, P.A., Mendonça Filho, J.G., da Silva, F.S., Flores, D., 2014. Solid bitumen occurrences in the Arruda sub-basin (Lusitanian 1042 Basin, Portugal): Petrographic features. International Journal of Coal Geology 131, 239–249. 1043 Goodarzi, F., Murchison, D., 1972. Optical properties of carbonized vitrinites. Fuel 51, 322–328. 1044 Goodarzi, F., 1984. Organic petrography of graptolite fragments from Turkey. Marine and Petroleum Geology 1, 202–210. 1045 Goodarzi, F., 1985a. Dispersion of optical properties of graptolite epiderms with increased maturity in early Paleozoic organic sediments. 1046 Fuel 64, 1735–1740. 1047 Goodarzi, F., 1985b. Reflected light microscopy of chitinozoan fragments. Marine and Petroleum Geology 2, 72–78. 1048 Goodarzi, F., Norford, B., 1985. Graptolites as indicators of the temperature histories of rocks. Journal of the Geological Society 142, 1049 1089–1099. 1050 Goodarzi, F., Snowdon, L., Gunther, P., Jenkins, W., 1985. Preliminary organic petrography of Palaeozoic rocks from the Grand Banks, 1051 Newfoundland. Marine and Petroleum Geology 2, 254–259. 1052 Goodarzi, F., Norford, B., 1987. Optical properties of graptolite epiderm—A review. Bulletin of Geological Society of Denmark 35, 141– 1053 147. 1054 Goodarzi, F., Gentzis, T., Feinstein, S., Snowdon, L., 1988. Effect of maceral subtypes and mineral matrix on measured reflectance of 1055 subbituminous coals and dispersed organic matter. International journal of coal geology 10, 383–398. 1056 Goodarzi, F., Norford, B., 1989. Variation of graptolite reflectance with depth of burial. International Journal of Coal Geology 11, 127– 1057 141. 1058 Goodarzi, F., Macqueen, R., 1990. Optical/compositional character of six bitumen species from Middle Devonian rocks of the Pine point 1059 Pb-Zn property, Northwest Territories, Canada. International Journal of Coal Geology 14, 197–216. 1060 Goodarzi, F., Fowler, M., Bustin, M., McKirdy, D., 1992a. Thermal maturity of Early Paleozoic sediments as determined by the optical 1061 properties of marine-derived organic matter—A review, Early Organic Evolution. Springer, pp. 279–295. 1062 Goodarzi, F., Gentzis, T., Harrison, C., Thorsteinsson, R., 1992b. The significance of graptolite reflectance in regional thermal maturity 1063 studies, Queen Elizabeth Islands, Arctic Canada. Organic Geochemistry 18, 347–357. 1064 Goodarzi, F., Gentzis, T., Harrison, J.C., Thorsteinsson, R., 1992c. Graptolite reflectance as indicator of thermal maturity of sedimentary 1065 rocks of Ordovician to Devonian age from the Queen Elizabeth Islands, Arctic Archipelago. Organic Geochemistry 18, 347–358. 1066 Goodarzi, F., Gentzis, T., Snowdon, L., Bustin, R., Feinstein, S., Labonte, M., 1993. Effect of mineral matrix and seam thickness on 1067 reflectance of vitrinite in high to low volatile bituminous coals: an enigma. Marine and Petroleum Geology 10, 162–171. 1068 Green, P.D., Johnson, C.A., Thomas, K.M., 1983. Applications of laser Raman microprobe spectroscopy to the characterization of coals 1069 and cokes. Fuel 62, 1013–1023. 1070 Gretener, P.E., 1982. Geothermics: using temperature in hydrocarbon exploration (revised). Am. Assoc. Pet. Geol., Short Course Note 1071 Series 17. 1072 Gupta, N.S., Briggs, D.E.G., Pancost, R.D., 2006. Molecular taphonomy of graptolites. Journal of the Geological Society 163, 897–900. 1073 Hackley, P.C., Cardott, B.J., 2016. Application of organic petrography in North American shale petroleum systems: A review. International 1074 Journal of Coal Geology 163, 8–51. 1075 Haeri-Ardakani, O., Sanei, H., Lavoie, D., Chen, Z., Jiang, C., 2015. Geochemical and petrographic characterization of the Upper 1076 Ordovician Utica Shale, southern Quebec, Canada. International Journal of Coal Geology 138, 83–94. 1077 Hao, J., Zhong, N., Luo, Q., Liu, D., Wu, J., Liu, A., 2019. Raman spectroscopy of graptolite periderm and its potential as an organic 1078 maturity indicator for the Lower Paleozoic in southwestern China. International Journal of Coal Geology 213, 103278. 1079 Henningsen, L.M., Jensen, C.H., Schovsbo. N.H., Nielsen. A.T., Pedersen, G.K., 2018. Shale fabric and organic nanoporosity in lower 1080 Palaeozoic shales, Bornholm, Denmark. Geological Survey of Denmark and Greenland Bulletin 41, 17–20. 1081 Henry, D.G., Jarvis, I., Gillmore, G., Stephenson, M., Emmings, J.F., 2018. Assessing low-maturity organic matter in shales using Raman 1082 spectroscopy: Effects of sample preparation and operating procedure. International Journal of Coal Geology 191, 135–151. 1083 Hickey, J.J., Henk, B., 2007. Lithofacies summary of the Barnett Shale, Mitchell 2 TP Sims well, Wise County, Texas. 1084 AAPG bulletin 91, 437–443. 1085 Hill, D.G., Nelson, C., 2000. Gas productive fractured shales: an overview and update. Gas Tips 6, 4–13. 1086 Hoffknecht, A., 1991. Mikropetrographische, organisch-geochemische, mikrothermometrische und mineralogische Unterschungen zur 1087 Bestimmung der Reife von Graptolithen-Periderm. Göttinger Arbeiten zur Geologie und Paläontologie 48, pp. 1–98. 1088 Hwang, R., Teerman, S., Carlson, R., 1998. Geochemical comparison of reservoir solid bitumens with diverse origins. Organic 1089 Geochemistry 29, 505–517. 1090 İnan, S., Goodarzi, F., Mumm, A.S., Arouri, K., Qathami, S., Ardakani, O.H., İnan, T., Tuwailib, A.A., 2016. The Silurian Qusaiba Hot 1091 Shales of Saudi Arabia: An integrated assessment of thermal maturity. International Journal of Coal Geology 159, 107–119. 1092 İnan, S., Al Badairy, H., İnan, T., Al Zahrani, A., 2018. Formation and occurrence of organic matter-hosted porosity in shales. International 1093 Journal of Coal Geology 199, 39–51. 1094 Jacob, H., 1989. Classification, structure, genesis and practical importance of natural solid oil bitumen (“migrabitumen”). International 1095 Journal of Coal Geology 11, 65–79. 1096 Jarvie, D.M., Hill, R.J., Ruble, T.E., Pollastro, R.M., 2007. Unconventional shale-gas systems: the Mississippian Barnett Shale of north- 1097 central Texas as one model for thermogenic shale-gas assessment. AAPG Bulletin 91, 475–499. 1098 Jarvie, D.M. 2012. Shale resource systems for oil and gas: Part 1 – Shale gas resource systems. AAPG Memoir 97, 69–87. 1099 Jehlička, J., Bény, C., 1992. Application of Raman microspectrometry in the study of structural changes in kerogens during 1100 regional metamorphism. Organic Geochemistry 18, 211–213. 1101 Jones, P.J., Stump, T.E., 1999. Depositional and tectonic setting of the lower Silurian hydrocarbon source rock facies, central Saudi Arabia. 1102 American Association of Petroleum Geologists Bulletin 83, 314–332. 1103 Jubb, A.M., Botterell, P.J., Birdwell, J.E., Burruss, R.C., Hackley, P.C., Valentine, B.J., Hatcherian, J.J., Wilson, S.A., 2018. High 1104 microscale variability in Raman thermal maturity estimates from shale organic matter. International Journal of Coal Geology 199, 1– 1105 9. 1106 Karcz, P., Janas, M., Dyrka, I., 2013. Polish shale gas deposits in relation to selected shale gas perspective areas of Central and Eastern 1107 Europe. Przeglad Geologiczny 61(7), 411–423. 1108 Karcz, P., Janas, M., 2016. Materia organiczna łupków kambru, ordowiku i syluru w baseniebałtycko-podlasko-lubelskim Polski. Przegląd 1109 Geologiczny 64, 995–999 (in Polish with English abstract). 1110 Kelemen, S.R., Fang, H.L., 2001. Maturity Trends in Raman Spectra from Kerogen and Coal. Energy & Fuels 15, 653–658. 1111 Kemp, A.E.S., Oliver, G.J.H., Baldwin, J.R., 1985. Low-grade metamorphism and accretion tectonics: Southern Uplands terrain, Scotland. 1112 Mineralogical Magazine 49, 335–344. 1113 Khatibi, S., Ostadhassan, M., Tuschel, D., Gentzis, T., Bubach, B., and Carvajal-Ortiz, H., 2018a. Raman spectroscopy to study thermal 1114 maturity and elastic modulus of kerogen. International Journal of Coal Geology, 185, 103–118. 1115 Khatibi, A., Ostadhassan, M., Tuschel, D., Gentzis, T., and Carvajal-Ortiz, H., 2018b. Evaluating molecular evolution of kerogen by Raman 1116 spectroscopy: Correlation with optical microscopy and Rock-Eval pyrolysis. Energies, 11 (6), 1–19. 1117 Khavari-Khorasani, G., 1975. The Properties and Structural Ordering of Some Fossil Bitumens. Ph.D. thesis, University of Newcastle 1118 upon Tyne, UK. 1119 Komorek, J., Morga, R., 2002. Relationship between the maximum and the random reflectance of vitrinite for coal from the Upper Silesian 1120 Coal Basin (Poland). Fuel 81, 969–971. 1121 Kozlowski, R., 1949. Les graptolites et quelque nouveaux groups d'animaux due Tremadocdela Pologne. Palaeontol. Pol. 31, 1–235. 1122 Krzesińska, M., Pusz, S., Smędowski, Ł., 2009. Characterization of the porous structure of cokes produced from the blends of three Polish 1123 bituminous coking coals. International Journal of Coal Geology 78, 169–176. 1124 Kurylowicz, L., Ozimic, S., McKirdy, D., Kantsler, A., Cook, A., 1976. Reservoir and source rock potential of the Larapinta Group, 1125 Amadeus Basin, central Australia. Australian Petroleum Exploration Association Journal 16, 49–65. 1126 Lavoie, D., Pinet, N., Bordeleau, G., Ardakani, O.H., Ladevèze, P., Duchesne, M.J., Rivard, C., Mort, A., Brake, V., Sanei, H., 2016. The 1127 Upper Ordovician black shales of southern Quebec (Canada) and their significance for naturally occurring hydrocarbons in shallow 1128 groundwater. International Journal of Coal Geology 158, 44–64. 1129 Lehr, J.H., Keeley, J. 2016. Alternative energy and shale gas encyclopedia. John Wiley & Sons, pp, 1–912. 1130 Leninger, A., 1975. Biochemistry. Worth Publishers, . 1131 Lin, R., Ritz, G.P., 1993. Reflectance FT-IR Microspectroscopy of Fossil Algae Contained in Organic-Rich Shales. Appl. Spectrosc. 47, 1132 265–271. 1133 Link, C., Bustin, R., Goodarzi, F., 1990. Petrology of graptolites and their utility as indices of thermal maturity in Lower Paleozoic strata 1134 in northern Yukon, Canada. International Journal of Coal Geology 15, 113–135. 1135 Littke, R., Krooss, B., Ufmann, A.K., Schultz, H.M., Horsfield, B., 2011. Unconventional gas resources in the Paleozoic of Central Europe. 1136 Oil Gas. Sci. Technol. Rév. IFP Energ. Nouv. 66, 953–977. 1137 Liu, D., Hou, X., Jiang, J., 1996. The composition and structure of graptolite-a micro-area analysis. Acta Mineralogica Sinica 16, 53–57 1138 (in Chinese with English abstract). 1139 Liu, D., Wang, l., Du, G., Hu, B., 2001. Study of the reflectance cross-plot of optical structure of graptolites in china. Geoscience 15, 321– 1140 325 (in Chinese with English abstract). 1141 Liu, D., Xiao, X., Tian, H., Min, Y., Zhou, Q., Cheng, P., Shen, J., 2013. Sample maturation calculated using Raman spectroscopic 1142 parameters for solid organics: Methodology and geological applications. Chinese Science Bulletin 58, 1285–1298. 1143 Loucks, R.G., Reed, R.M., Ruppel, S.C., Jarvie, D.M., 2009. Morphology, genesis, and distribution of nanometer-scale pores in siliceous 1144 mudstones of the Mississippian Barnett Shale. Journal of Sedimentary Research 79, 848–861. 1145 Loucks, R.G., Reed, R.M., Ruppel, S.C., Hammes, U., 2012. Spectrum of pore types and networks in mudrocks and a descriptive 1146 classification for matrix-related mudrock pores. AAPG Bulletin 96, 1071–1098. 1147 Loydell, D.K., 1998. Early Silurian sea-level changes. Geological Magazine 135, 447–471. 1148 Loydell, D.K., 2011. Graptolite biozone correlation charts. Geological Magazine 149, 124–132. 1149 Luo, Q., Zhong, N., Dai, N., Zhang, W., 2016. Graptolite-derived organic matter in the Wufeng–Longmaxi Formations (Upper Ordovician– 1150 Lower Silurian) of southeastern Chongqing, China: Implications for gas shale evaluation. International Journal of Coal Geology 153, 1151 87–98. 1152 Luo, Q., Hao, J., Skovsted, C.B., Luo, P., Khan, I., Wu, J., Zhong, N., 2017. The organic petrology of graptolites and maturity assessment 1153 of the Wufeng–Longmaxi Formations from Chongqing, China: Insights from reflectance cross-plot analysis. International Journal of 1154 Coal Geology 183, 161–173. 1155 Luo, Q., Hao, J., Skovsted, C.B., Xu, Y., Liu, Y., Wu, J., Zhang, S., Wang, W., 2018. Optical characteristics of graptolite-bearing sediments 1156 and its implication for thermal maturity assessment. International Journal of Coal Geology 195, 386–401. 1157 Luo, Q., Hao, J., Li, K., Xu, Y., Wang, X., Wang, H., Luan, J., Hu, K., Li, T., Zhong, N., 2019. A new parameter for the thermal maturity 1158 assessment in the palaeozoic sediments: a restudy on the optical characteristics of the graptolite. Acta Geologica Sinica (Chinese 1159 edition), 93(9), 2362–2371 (in Chinese with English abstract). 1160 Lüning, S., Craig, J., Loydell, D.K., Štorch, P., Fitches, B., 2000. Lower Silurian `hot shales' in North Africa and Arabia: regional 1161 distribution and depositional model. Earth Science Reviews 49, 121–200. 1162 Lüning, S., Kolonic, S., Loydell, D.K., Craig, J., 2003. Reconstruction of the original organic richness in weathered Silurian shale outcrops 1163 (Murzuq and Kufra basins, southern Libya). GeoArabia 8, 299–308. 1164 Lüning, S., Shahin, Y., Loydell, D., Al-Rabi, H., Masri, A., Tarawneh, B., Kolonic, S., 2005. Anatomy of a world-class source rock: 1165 Distribution and depositional model of Silurian organic-rich shales in Jordan and implications for hydrocarbon potential. AAPG 1166 bulletin 89, 1397–1427. 1167 Lünsdorf, N.K., 2016. Raman spectroscopy of dispersed vitrinite—Methodical aspects and correlation with reflectance. International 1168 Journal of Coal Geology 153, 75–86. 1169 Ma, X., Xie, J., 2018. The progress and prospects of shale gas exploration and development in southern Sichuan Basin, SW China. 1170 Petroleum Exploration and Development 45, 172–182. 1171 Ma, Y., Zhong, N., Cheng, L., Pan, Z., Dai, N., Zhang, Y., Yang, L., 2016. Pore structure of the graptolite-derived OM in the Longmaxi 1172 Shale, southeastern Upper Yangtze Region, China. Marine and Petroleum Geology 72, 1–11. 1173 Makarov, K.K., Bazhenova, T.K., 1981. Organic Geochemistry of the Paleozoic and Pre-Paleozoic of the Siberian Platform and the 1174 Perspectives of Oil and Gas. Nedra Publ. House, Leningrad, 209 pp. (in Russian) 1175 Malinconico, M.L., 1992. Graptolite reflectance in the prehnite-pumpellyite zone of Northern Maine, USA. Organic Geochemistry 18, 1176 263–271. 1177 Malinconico, M.L., 1993. Reflectance cross-plot analysis of graptolites from the anchi-metamorphic region of northern Maine, USA. 1178 Organic Geochemistry 20, 197–207. 1179 Marshall, R., Murchison, D., 1971. Dispersion of the optical properties of carbonized vitrinites. Fuel 50, 4–22. 1180 Mastalerz, M., Bustin, R.M., 1993a. Electron microprobe and micro-FTIR analyses applied to maceral chemistry. International Journal of 1181 Coal Geology 24, 333–345. 1182 Mastalerz, M., Bustin, R.M., 1993b. Variation in maceral chemistry within and between coals of varying rank: An electron microprobe 1183 and micro-Fourier transform infra-red investigation. Journal of Microscopy 171, 153–166. 1184 Mastalerz, M., Bustin, R.M., 1995. Application of reflectance micro-Fourier transform infrared spectrometry in studying coal macerals: 1185 comparison with other Fourier transform infrared techniques. Fuel 74, 536–542. 1186 Mastalerz, M., Bustin, R.M., 1996. Application of reflectance micro-Fourier transform infrared analysis to the study of coal macerals: an 1187 example from the late to early coals of the Mist Mountain Formation, British Columbia, Canada. International 1188 Journal of Coal Geology 32, 55–67. 1189 Mastalerz, M., Bustin, R.M., 1997. Variation in the chemistry of macerals in coals of the Mist Mountain Formation, Elk Valley coalfield, 1190 British Columbia, Canada. International Journal of Coal Geology 33, 43–59. 1191 Mastalerz, M., Schimmelmann, A., Drobniak, A., Chen, Y., 2013. Porosity of Devonian and Mississippian across a 1192 maturation gradient: insights from organic petrology, gas adsorption, and mercury intrusion. AAPG Bulletin 97, 1621–1643. 1193 Mastalerz, M., Drobniak, A., Stankiewicz, A.B., 2018. Origin, properties, and implications of solid bitumen in source-rock reservoirs: A 1194 review. International Journal of Coal Geology 195, 14–36. 1195 Milliken, K.L., Rudnicki, M., Awwiller, D.N., Zhang, T., 2013. Organic matter–hosted pore system, (Devonian), 1196 . AAPG Bulletin 97, 177–200. 1197 Modliński, Z., Szymański, B., 1997. The Ordovician lithostratigraphy of the Peribaltic Depression (NE Poland). Geological Quarterly 41, 1198 273–288. 1199 Modliński, Z., Szymański, B., Teller, L., 2006. Litostratygrafia syluru polskiej części obniżenia perybałtyckiego—część lądowa i morska 1200 (N Polska). Przegląd Geologiczny 54, 787–796 (in Polish with English abstract). 1201 Modliński, Z., Szymański, B., 2008. Litostratygrafia ordowiku w obniżeniu podlaskim i w podłożu niecki płocko-warszawskiej 1202 (wschodnia Polska). Biul. Państw. Inst. Geol. 430, 79–112 (in Polish with English abstract). 1203 Moore, R.C., 1955. Treatise on Invertebrate Paleontology, Part V, Graptolithina. Kansas, Geol. Soc. Am. University of Kansas Press. 1204 Morga, R., Kamińska, M., 2018. The chemical composition of graptolite periderm in the gas shales from the Baltic Basin of Poland. 1205 International Journal of Coal Geology 199, 10–18. 1206 Morga, R., Pawlyta, M., 2018. Microstructure of graptolite periderm in Silurian gas shales of Northern Poland. International Journal of 1207 Coal Geology 189, 1–7. 1208 Mumm, A.S., İnan, S., 2016. Microscale Organic Maturity determination of Graptolites using Raman Spectroscopy. International Journal 1209 of Coal Geology 162, 96–107. 1210 Nielsen, A.T., Schovsbo, N.H., 2007. Cambrian to basal Ordovician lithostra-tigraphy in southern Scandinavia. Bulletin of the Geological 1211 Society of Denmark 53, 47–92. 1212 Nielsen, A.T., Schovsbo, N.H., 2011. The Lower Cambrian of Scandinavia: Depositional environment, sequence stratigraphy and 1213 palaeogeography. Earth-Science Reviews 107, 207–310. 1214 Nielsen, A.T., Schovsbo, N.H., 2015. The regressive Early-Mid Cambrian 'Hawke Bay Event' in Baltoscandia: Epeirogenic uplift in concert 1215 with eustasy. Earth-Science Reviews 151, 288–350. 1216 Oliver, G.J.H., 1988. Arenig to Wenlock regional metamorphism in the paratectonic Caledonides of the British Isles: A review. In: The 1217 Caledonian-Appalachian Orogenesis (Harris, A.L. and Fettes, D.J., eds.), Spec. Publ. Geol. Soc. London 38, 347–363. 1218 Passey, Q.R., Bohacs, K., Esch, W.L., Klimentidis, R., Sinha, S., 2010. From oil-prone source rock to gas-producing shale reservoir– 1219 geologic and petrophysical characterization of unconventional shale-gas reservoirs, Chinese Petroleum Society/Society of Petroleum 1220 Engineers International Oil & Gas Conference and Exhibition, Beijing, China, p. 29. 1221 Pawlyta, M., 2013. Transmission electron microscope studies on carbon nanostructured materials. Arch. Mater. Sci. Eng. 63, 58–67. 1222 Peters, K., 1986. Guidelines for evaluating petroleum source rock using programmed pyrolysis. AAPG Bulletin 70, 318–329. 1223 Peters, K.E., Cassa, M.R., 1994. Applied Source Rock Geochemistry, in: Magoon, L.B., Dow, W.G. (Eds.), The Petroleum System—From 1224 Source to Trap. American Association of Petroleum Geologists Bulletin, pp. 93–120. 1225 Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The biomarker guide. Cambridge University Press, Cambridge. 1226 Petersen, H.I., Schovsbo, N.H., Nielsen, A.T., 2013. Reflectance measurements of zooclasts and solid bitumen in Lower Palaeozoic shales, 1227 southern Scandinavia: Correlation to vitrinite reflectance. International Journal of Coal Geology 114, 1–18. 1228 PIG (Państwowy Instytut Geologiczny), 2012. Ocena zasobów wydobywalnych gazu ziemnego i ropy naftowej w formacjach łupkowych 1229 dolnego paleozoiku w Polsce (Basen Bałtycko-Podlasko-Lubelski). Raport pierwszy. 1230 Podhalańska, T., 2013. Graptolity–narzędzie stratygraficzne w rozpoznaniu stref perspektywicznych dla występowania 1231 niekonwencjonalnych złóż węglowodorów. Przegląd Geologiczny 460–467 (in Polish with English abstract). 1232 Podhalańska, T., Waksmundzka, M., Becker, A., Roszkowska-Remin, J., Dyrka, I., Feldman-Olszewska, A., Głuszyński, A., Grotek, I., 1233 Janas, M., Karcz, P., Nowak, G., Pacześna, J., Roman, M., Sikorska-Jaworowska, M., Kuberska, M., Kozłowska, A., Sobień, K., 1234 2016a. Strefy perspektywiczne występowania niekonwencjonalnych złóż węglowodorów w kambryjskich, ordowickich, sylurskich i 1235 karbońskich kompleksach skalnych Polski–integracja wyników badań. Przegląd Geologiczny 64, 1008–1021 (in Polish with English 1236 abstract). 1237 Podhalańska, T., Waksmundzka, M., Becker, A., Roszkowska-Remin, J., 2016b. Rozpoznanie stref perspektywicznych występowania 1238 niekonwencjonalnych złóż węglowodorów w Polsce–nowe wyniki oraz dalsze kierunki badań. Przegląd Geologiczny 64, 953–962 1239 (in Polish with English abstract). 1240 Pool, W., Geluk, M., Abels, J., Tiley, G., 2012. Assessment of an unusual European Shale Gas play – The Cambro–Ordovician Alum Shale, 1241 southern Sweden: Proceedings of the Society of Petroleum Engineers/European Association of Geoscientists and Engineers 1242 Unconventional Resources Conference, Vienna, Austria, 20–22 March, 2012. SPE 152339-MS. 1243 Poprawa, P., 2010. Potencjał występowania złóż gazu ziemnego w łupkach dolnego paleozoiku w basenie bałtyckim i lubelsko-podlaskim. 1244 Przegląd Geologiczny 58, 226–249 (in Polish with English abstract). 1245 Qiu, Z., Zou, C., Li, X., Wang, H., Dong, D., Lu, B., Zhou, S., Shi, Z., Feng, Z., Zhang, M., 2018. Discussion on the contribution of 1246 graptolite to organic enrichment and gas shale reservoir: A case study of the Wufeng–Longmaxi shales in South China. Journal of 1247 Natural Gas Geoscience 3, 147–156. 1248 Radke, M., 1988. Application of aromatic compounds as maturity indicators in source rocks and crude oils. Marine and Petroleum Geology 1249 5, 224–236. 1250 Radke, M., Welte, D., 1983. The methylphenanthrene index (MPI): a maturity parameter based on aromatic hydrocarbons. In: Bjorøy, M. 1251 (Ed.), Advances in organic geochemistry, 1981. Wiley, Chichester, pp. 504–512. 1252 Radke, M., Welte, D., Willsch, H., 1986. Maturity parameters based on aromatic hydrocarbons: Influence of the organic matter type. 1253 Organic Geochemistry 10, 51–63. 1254 Rantitsch, G., 1995. Coalification and graphitization of graptolites in the anchizone and lower epizone. International Journal of Coal 1255 Geology 27, 1–22. 1256 Rantitsch, G., 1997. Thermal history of the Carnic Alps (Southern Alps, Austria) and its paleogeographic implications. Tectonophysics 1257 272, 213–232. 1258 Reyes, J., Jiang, C., Lavoie, D., Armstrong, D.K., Milovic, M., Robinson, R., 2018. Organic petrographic analysis of artificially matured 1259 chitinozoan- and graptolite-rich Upper Ordovician shale from Hudson Bay Basin, Canada. International Journal of Coal Geology 199, 1260 138–151. 1261 Riediger, C., Goodarzi, F., Macqueen, R., 1989. Graptolites as indicators of regional maturity in lower Paleozoic sediments, Selwyn Basin, 1262 Yukon and Northwest Territories, Canada. Canadian Journal of Earth Sciences 26, 2003–2015. 1263 Romankevich, E.A., 1984. Geochemistry of organic matter in the ocean. Springer-Verlag, Berlin. 1264 Romero-Sarmiento, M.F., Rouzaud, J.N., Bernard, S., Deldicque, D., Thomas, M., Littke, R., 2014. Evolution of Barnett Shale organic 1265 carbon structure and nanostructure with increasing maturation. Organic Geochemistry 71, 7–16. 1266 Ross, D.J., Bustin, R.M., 2007. Shale gas potential of the lower jurassic gordondale member, northeastern British Columbia, Canada. 1267 Bulletin of Canadian Petroleum Geology 55, 51–75. 1268 Ross, D.J., Bustin, R.M., 2009. The importance of shale composition and pore structure upon gas storage potential of shale gas reservoirs. 1269 Marine and Petroleum Geology 26, 916–927. 1270 Sanei, H., Petersen, H., Schovsbo, N., Jiang, C., Goodsite, M.E., 2014. Petrographic and geochemical composition of kerogen in the 1271 Furongian (U. Cambrian) Alum Shale, central Sweden: Reflections on the petroleum generation potential. International Journal of 1272 Coal Geology 132, 158–169. 1273 Sauerer, B., Craddock, P.R., AlJohani, M.D., Alsamadony, K.L., Abdallah, W., 2017. Fast and accurate shale maturity determination by 1274 Raman spectroscopy measurement with minimal sample preparation. International Journal of Coal Geology 173, 150–157. 1275 Schmidt, J.S., Menezes, T.R., Souza, I.V.A.F., Spigolon, A.L.D., Pestilho, A.L.S., Coutinho, L.F.C., 2019. Comments on empirical 1276 conversion of solid bitumen reflectance for thermal maturity evaluation. International Journal of Coal Geology 201, 44–50. 1277 Schoenherr, J., Littke, R., Urai, J.L., Kukla, P.A., Rawahi, Z., 2007. Polyphase thermal evolution in the Infra-Cambrian Ara Group (South 1278 Oman Salt Basin) as deduced by maturity of solid reservoir bitumen. Organic Geochemistry 38, 1293–1318. 1279 Schovsbo, N.H., 2003. Geochemical composition and provenance of Lower Palaeozoic shales deposited at the margins of Baltica. Bulletin 1280 of the Geological Society of Denmark 50, 11–27. 1281 Schovsbo, N.H., Nielsen, A.T., Klitten, K., Mathiesen, A., Rasmussen, P., 2011. Shale gas investigations in Denmark: Lower Palaeozoic 1282 shales on Bornholm. Geological Survey of Denmark and Greenland Bulletin 23, 9–12. 1283 Schovsbo, N.H., Nielsen, A.T., Gautier, D.L., 2014. The Lower Palaeozoic shale gas play in Denmark. Geological Survey of Denmark and 1284 Greenland Bulletin 31, 19–22. 1285 Schovsbo, N.H., Nielsen, A.T., 2017. Generation and origin of natural gas in Lower Palaeozoic shales from southern Sweden. Geological 1286 Survey of Denmark and Greenland Bulletin 38, 37–40. 1287 Schovsbo, N.H., Nielsen A.T., Harstad, A.O., Bruton, D.L., 2018. Stratigraphy and geochemical composition of the Cambrian Alum Shale 1288 Formation in the Porsgrunn core, Skien-Langesund district, southern Norway. Bulletin of the Geological Society of Denmark 66, 1– 1289 20. 1290 Schovsbo, N.H., Jakobsen, F., 2019. Review of hydrocarbon potential in East Denmark following 30 years of exploration activities. 1291 Geological Survey of Denmark and Greenland Bulletin 43, e2019430105. https://doi.org/10.34194/GEUSB-201943-01-05. 1292 Schulz, H.M., Biermann, S., van Berk, W., Krüger, M., Straaten, N., Bechtel, A., Wirth, R., Lüders, V., Schovsbo, N.H., Crabtree, S. 2015. 1293 From shale oil to biogenic shale gas: retracing organic-inorganic interactions in the Alum Shale (Furongian-Lower Ordovician) in 1294 southern Sweden. AAPG Bulletin 99, 927–956. 1295 Soua, M., 2014. Paleozoic oil/gas shale reservoirs in southern Tunisia: An overview. Journal of African Earth Sciences 100, 450–492. 1296 Stach, E., Mackowski, M.T., Teichmuller, M., Taylor, G.H., Chandra, D., Teichmuller, R., 1982. Stach's Textbook of Coal Petrology. 1297 Gebruder Borntraeger, Berlin. 1298 Suárez-Ruiz, I., Flores, D., Mendonça Filho, J.G., Hackley, P.C., 2012. Review and update of the applications of organic petrology: Part 1299 1, geological applications. International Journal of Coal Geology 99, 54–112. 1300 Suchý, V., Sýkorová, I., Stejskal, M., Šafanda, J., Machovič, V., Novotná, M., 2002. Dispersed organic matter from Silurian shales of the 1301 Barrandian Basin, Czech Republic: optical properties, chemical composition and thermal maturity. International journal of coal 1302 geology 53, 1–25. 1303 Suchý, V., Šafanda, J., Sýkorová, I., Stejskal, M., Machovič, V., Melka, K., 2004. Contact metamorphism of Silurian black shales by a 1304 basalt sill: geological evidence and thermal modelling in the Barrandian Basin. Bull. Geosci. 79, 133–145. 1305 Suchý, V., Sandler, A., Slobodník, M., Sýkorová, I., Filip, J., Melka, K., Zeman, A., 2015. Diagenesis to very low-grade metamorphism in 1306 lower Palaeozoic sediments: A case study from deep borehole Tobolka 1, the Barrandian Basin, Czech Republic. International journal 1307 of coal geology 140, 41–62. 1308 Synnott, D.P., Dewing, K., Ardakani, O.H., Obermajer, M., 2018. Correlation of zooclast reflectance with Rock-Eval Tmax values within 1309 Upper Ordovician Cape Phillips Formation, a potential petroleum source rock from the Canadian Arctic Islands. Fuel 227, 165–176. 1310 Szymański, B. 2008. Zapis litologiczny i mikrofacjalny osadów euksynicznych kambru górnego i tremadoku obniżenia bałtyckiego 1311 (północna Polska). Biuletyn PIG 430, 113–154 (in Polish with English abstract). 1312 Tasch, P., 1980. Paleobiology of the Invertebrates, Data Retrieval from the Fossil Record. Wiley, New York. 1313 Taylor, G.H., Teichmqller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic Petrology. Gebrqder Borntrager, Berlin. 1314 Teichmüller, M., 1978. Nachweis von graptolithen-periderm in geschieferten gesteinen mit hilfe kohlenpetrologischer methoden. Neues 1315 Jahrbuch für Geologie und Paläontologie, Monatshefte 7, 430–447. 1316 Teichmüller, M., Teichmüller, R., Bartenstein, H., 1979. Inkohlung und Erdgas in Nordwestdeutschland. Eine Inkohlungskarte der 1317 oberfläche des oberkarbons. Fortschritte in der Geologie von Rheinland und Westfalen 27, 137–170. 1318 Teichmüller, M., 1982. Origin of the petrographic constituents of coal. In: Stach, E., Mackowsky, M., Teichmuller, M., Taylor, G.H., 1319 Chandra, D., Teichmuller, R. (Eds.), Stach's Textbook of coal petrology. Gebruder Borntraeger Berlin, pp. 219–294. 1320 Teichmüller, M., 1987. Organic material and very low-grade metamorphism. In: Frey, M. (ed.), Low Temperature Metamorphism. Blackie, 1321 London and Glasgow, pp. 114–161. 1322 Tian, H., Pan, L., Xiao, X., Wilkins, R.W., Meng, Z., Huang, B., 2013. A preliminary study on the pore characterization of Lower Silurian

1323 black shales in the Chuandong Thrust Fold Belt, southwestern China using low pressure N2 adsorption and FE-SEM methods. Marine 1324 and Petroleum Geology 48, 8–19. 1325 Towe, K.M., Urbanek, A., 1972. Collagen-like Structures in Ordovician Graptolite Periderm. Nature 237, 443–445. 1326 Trettin, H.P., 1989. The Arctic Islands. In Bally, A.W., Palmer, A.R. (Eds.), The Geology of North America. Geol. Soc. America., 1, pp. 1327 349–370. 1328 Trettin, H.P., 1990. Geotectonic correlation chart. In Bally, A.W., Palmer, A.R., (Eds). Innuitian Orogen and Arctic Platform, Canada and 1329 Greenland. Geol. Surv. Can. 3, Sheet I. 1330 Tricker, P.M., Marshall, J.E., Badman, T.D., 1992. Chitinozoan reflectance: a Lower Palaeozoic thermal maturity indicator. Marine and 1331 Petroleum Geology 9, 302–307. 1332 Urbanek, A., Mierzejewski, P., 1986. A possible new pattern of cortical deposit in Tremadoc dendroid graptolites from chert nodules. 1333 Geological Society, London, Special Publications 20, 13–19. 1334 van Krevelen, D., 1961. Coal. Elsevier, Amsterdam. 1335 Varol, O.N., Demirel, I.H., Rickards, R.B., Gunay, Y., 2006. Source rock characteristics and biostratigraphy of the Lower Silurian 1336 (Telychian) organic-rich shales at Akyaka, central Taurus region, Turkey. Marine and Petroleum Geology 23 (9-10), 901–911. 1337 Wang, F., He, P., Cheng, D., Hao, S., 1994. The thermal maturity evaluation of the lower palaeozoic high-postmature hydrocarbon source 1338 rocks. Natural Gas Geoscience 26, 1–14 (in Chinese). 1339 Wang, D., Philp, R.P., 1997. Geochemical study of potential source rocks and crude oils in the Anadarko basin, Oklahoma. American 1340 Association of Petroleum Geologists Bulletin 81(2), 249–275. 1341 Wang, Q., Qian, M., Jiang, Q., Yang, Y., Tenger, B., 2017a. A study on hydrocarbon generation capacity of graptolite in marine hydrocarbon 1342 source rocks in Southern China. Rock and Mineral Analysis 36, 258–264 (in Chinese with English abstract). 1343 Wang, S., Tang, Y., Schobert, H.H., Jiang, D., Guo, X., Huang, F., Guo, Y., Su, Y., 2014. Chemical compositional and structural 1344 characteristics of Late bark coals from Southern China. Fuel 126, 116–121. 1345 Wang, S., Liu, S., Sun, Y., Jiang, D., Zhang, X., 2017b. Investigation of coal components of Late Permian different ranks bark coal using 1346 AFM and Micro-FTIR. Fuel 187, 51–57. 1347 Wang, X., Hoffknecht, A., Xiao, J., Chen, S., Li, Z., Bainer Brocke, R., Erdtmann, B.D., 1993. Graptolite, chitinozoan and scolecodont 1348 reflectances and their use as indicators of thermal maturity. Acta Geologica Sinica (English Edition) 6, 93–105. 1349 Wang, Y., Qiu, N., Borjigin, T., Shen, B., Xie, X., Ma, Z., Lu, C., Yang, Y., Yang, L., Cheng, L., Fang, G., Cui, Y., 2019a. Integrated 1350 assessment of thermal maturity of the Upper Ordovician–Lower Silurian Wufeng–Longmaxi shale in Sichuan Basin, China. Marine 1351 and Petroleum Geology 100, 447–465. 1352 Wang, Z., Shi, B., Wen, Z., Tong, X., Song, C., He, Z., Liu, X., 2019b. Shale oil and gas exploration potential in the Tanezzuft Formation, 1353 Ghadames Basin, North Africa. Journal of African Earth Sciences 153, 83–90. 1354 Ward, C.R., Gurba, L.W., 1999. Chemical composition of macerals in bituminous coals of the Gunnedah Basin, Australia, using electron 1355 microprobe analysis techniques. International Journal of Coal Geology 39, 279–300. 1356 Ward, C.R., Li, Z., Gurba, L.W., 2005. Variations in coal maceral chemistry with rank advance in the German Creek and Moranbah Coal 1357 Measures of the Bowen Basin, Australia, using electron microprobe techniques. International Journal of Coal Geology 63, 117–129. 1358 Ward, C.R., Li, Z., Gurba, L.W., 2007. Variations in elemental composition of macerals with vitrinite reflectance and organic sulphur in 1359 the Greta Coal Measures, New South Wales, Australia. International Journal of Coal Geology 69, 205–219. 1360 Ward, C.R., Li, Z., Gurba, L.W., 2008. Comparison of elemental composition of macerals determined by electron microprobe to whole- 1361 coal ultimate analysis data. International Journal of Coal Geology 75, 157–165. 1362 Więcław, D., Kotarba, M., Kosakowski, P., Kowalski, A., Grotek, I., 2010. Habitat and hydrocarbon potential of the lower Paleozoic source 1363 rocks in the Polish part of the Baltic region. Geological Quarterly 54, 159–182. 1364 Wilkins, R.W.T., Boudou, R., Sherwood, N., Xiao, X., 2014. Thermal maturity evaluation from inertinites by Raman spectroscopy: The 1365 ‘RaMM’ technique. International Journal of Coal Geology 128-129, 143–152. 1366 Wilkins, R.W.T., Wang, M., Gan, H., Li, Z., 2015. A RaMM study of thermal maturity of dispersed organic matter in marine source rocks. 1367 International Journal of Coal Geology 150-151, 252–264. 1368 Wilkins, R.W.T., Sherwood, N., Li, Z., 2018. RaMM (Raman maturity method) study of samples used in an interlaboratory exercise on a 1369 standard test method for determination of vitrinite reflectance on dispersed organic matter in rocks. Marine and Petroleum Geology 1370 91, 236–250. 1371 Williams, S.H., Burden, E.T., Mukhopadhyay, P., 1998. Thermal maturity and burial history of Paleozoic rocks in western Newfoundland. 1372 Canadian Journal of Earth Sciences 35, 1307–1322. 1373 Wolvers, H.M., Maletz, J., 2016. The benthic graptolite Sphenoecium mesocambricus (Öpik, 1933) from the Middle Cambrian of Krekling, 1374 Oslo Region, Norway. Norwegian Journal of Geology 96, 311–318. 1375 Wopenka, B., Pasteris, J.D., 1993. Structural characterization of kerogens to granulite-facies graphite: Applicability of Raman microprobe 1376 spectroscopy. American Mineralogist 78, 533–557. 1377 Xiao, X., Liu, D., Fu, J., Wilkins, R.W., 1997. Marine vitrinite-an important hydrocarbon source matter in marine source rocks. Acta 1378 Petrolei Sinica 18, 44–48 (in Chinese with English abstract). 1379 Xiao, X., Wilkins, R., Liu, D., Liu, Z., Fu, J., 2000. Investigation of thermal maturity of lower Palaeozoic hydrocarbon source rocks by 1380 means of vitrinite-like maceral reflectance–a Tarim Basin case study. Organic Geochemistry 31, 1041–1052. 1381 Yang, C., Hesse, R., 1993. Diagenesis and anchimetamorphism in an overthrust belt, external domain of the Taconian Orogen, southern 1382 Canadian Appalachians-II. Paleogeothermal gradients derived from maturation of different types of organic matter. Organic 1383 Geochemistry 20(3), 381–403. 1384 Yang, S., Schulz, H.M., Schovsbo, N.H., Bojesen-Koefoed, J.A., 2017. Oil-source-rock correlation of the Lower Paleozoic petroleum 1385 system in the Baltic Basin (northern Europe). AAPG Bulletin 101, 1971–1993. 1386 Yang, Y., 2016. Application of bitumen and graptolite reflectance in the Silurian Longmaxi shale, southeastern Sichuan Basin. Petroleum 1387 Geology & Experiment 38, 466–472 (in Chinese with English abstract). 1388 Zerda, T.W., John, A., Chmura, K., 1981. Raman studies of coals. Fuel 60, 375–378. 1389 Zhang, J., Li, X., Wei, Q., Gao, W., Liang, W., Wang, Z., Wang, F., 2017. Quantitative characterization of pore-fracture system of organic- 1390 rich marine-continental shale reservoirs: A case study of the Upper Permian Longtan Formation, Southern Sichuan Basin, China. Fuel 1391 200, 272–281. 1392 Zhong, N., Qin, Y., 1995. Organic Petrology of Carbonate Rocks: Characteristics, Origin and Evolution of Macerals with Respects to 1393 Hydrocarbon Generation. Science Press, Beijing (in Chinese). 1394 Zhu, Y., Han, Z., Wei, K., Wang, Y., Chen, S., 2015. Organic Nanopores of Longmaxi and Qiongzhusi Formations in the Upper Yangtze: 1395 biological precursor and pore network. Natural Gas Geoscience 26, 1507–1515 (in Chinese with English abstract). 1396 Zijp, M., Nelskamp, S., Schovsbo, N.H., Tougaard, L., Bocin-Dumitriu, A., 2017. Resource estimation of Eighty-Two European shale 1397 formations. Unconventional Resources Technology Conference (URTeC) 2017 in Austin, Texas. Extended abstract 1-8. 1398 https://doi.org/10.15530-urtec-2017-2686270. 1399 Zou, C., Dong, D., Wang, S., Li, J., Li, X., Wang, Y., Li, D., Cheng, K., 2010. Geological characteristics and resource potential of shale 1400 gas in China. Petroleum Exploration and Development 37, 641–653. 1401 Zou, C., Yang, Z., Tao, S., Li, W., Wu, S., Hou, L., Zhu, R., Yuan, X., Wang, L., Gao, X., Jia, J., Guo, Q., Bai, B., 2012. Nano-hydrocarbon 1402 and the accumulation in coexisting source and reservoir. Petroleum Exploration and Development 39, 13–26 (in Chinese with English 1403 abstract). 1404 Zou, C., Dong, D., Wang, Y., Li, X., Huang, J., Wang, S., Guan, Q., Zhang, C., Wang, H., Liu, H., Bai, W., Liang, F., Lin, W., Zhao, Q., 1405 Liu, D., Yang, Z., Liang, P., Sun, S., Qiu, Z., 2015. Shale gas in China: Characteristics, challenges and prospects (I). Petroleum 1406 Exploration and Development 42, 753–767. 1407 Zou, C., Dong, D., Wang, Y., Li, X., Huang, J., Wang, S., Guan, Q., Zhang, C., Wang, H., Liu, H., Bai, W., Liang, F., Lin, W., Zhao, Q., 1408 Liu, D., Yang, Z., Liang, P., Sun, S., Qiu, Z., 2016. Shale gas in China: Characteristics, challenges and prospects (II). Petroleum 1409 Exploration and Development 43, 182–196. 1410