You are on page 1of 42

Elsevier Editorial System(tm) for Quaternary

Geochronology
Manuscript Draft

Manuscript Number: QUAGEO-D-15-00047R2

Title: CALCULATION OF THE RESERVOIR AGE FROM ORGANIC AND CARBONATE


FRACTIONS OF SEDIMENTS IN THE GULF OF CARIACO (CARIBBEAN SEA)

Article Type: Research Paper

Keywords: late Holocene, Reservoir effect, 14C dating AMS, Gulf of


Cariaco, LOI, Rock-Eval Pyrolysis.

Corresponding Author: Mrs. Iliana Aguilar,

Corresponding Author's Institution: Université Savoie-Mont-Blanc

First Author: Iliana Aguilar, Phd student

Order of Authors: Iliana Aguilar, Phd student; Pierre Sabatier; Christian


Beck; Franck Audemard; Christian Crouzet; Franco Urbani; Corina Campos

Abstract: A set of 24 AMS 14C measurements were performed on different


sediment fractions from short gravity cores taken the Gulf of Cariaco
(Northeastern Venezuela), in order to construct a detailed sedimentary
archive for the last centuries. There, a local reservoir effect was
expected because of: i) strong upwelling from the neighbouring Cariaco
Trough and anoxia, ii) specific geomorphological setting related to
active fault with seepages and diapirism. Measurements was thus performed
on different components: sediment bulk organic fraction (SBOF), pteropods
shells (surface water), bivalve shells (bottom), and plant fragments.
Based on comparisons of the different results, we discuss different age
corrections and calibrations, which lead us to consider a localized
reservoir correction (ΔR) negligible for bivalve and pteropod shells,
showing no reservoir age for carbonate and no water stratification. At
the difference, a local 633 ± 64 yr ΔR appears recorded in the organic
fraction of sedimentation and its origin is primary related to the
upwelling mechanism.
Cover Letter

COVER LETTER

Iliana Aguilar
ISTerre
UMR 5275 CNRS – Université Savoie Mont Blanc
73376 Le Bourget du Lac CEDEX, France
iliana.aguilar-ramos@univ-savoie.fr

Dear Editor,

Refering to your decision letter from October 1st 2015, we would like to submit a revised
version of our previous manuscript “Calculation of the reservoir age from organic and
carbonate fractions of sediments in the gulf of Cariaco (Caribbean Sea)” for publication
in Quaternary Geochronology.

First of all, we would like to thank you and the reviewer for their constructive comments.
Through them we have improved the article and clarified some aspects. Briefly, we have
carefully considered all Reviewer comments and modified the pertinent sections taking into
account the strict space requirements of the journal.

Please find below the response to each of the individual comments made by the Reviewer.
Any associated modification will be indicated with the expression "Modified in the text» with
the reference line in the annotated manuscript.

The writing and grammar changes were not specified in this letter but were systematically
incorporated in the annotated version. A colleague, fluent English speaker/writer, has
contributed to the corrected version of this manuscript.

We have indicated the all changes in the annotated version of the revised manuscript in
red.Minor modifications were made in Figures 1, 5 and Figure 6

*********************
*Detailed Response to Reviewers

REVISION

Iliana Aguilar
ISTerre
UMR 5275 CNRS – Université Savoie Mont Blanc
73376 Le Bourget du Lac CEDEX, France
iliana.aguilar-ramos@univ-savoie.fr

Dear Editor,

The writing and grammar changes were not specified in this letter but were systematically
incorporated in the annotated version. A colleague, fluent English speaker/writer, has
contributed to the corrected version of this manuscript.

We have indicated all the changes in the annotated version of the revised manuscript in
red. Minor modifications were made in Figures 7 and 8 and in the table 1 and 2 in relation to
ΔR error calculation suggested by the reviewer (see below).

*********************
Reviewer Comments:

1- I have made a few comments on the pdf that need to be addressed but these are
minor. The authors have not made any attempt to improve the quality of the grammar. I
realize that this is difficult but there is the possibility that some statements could be
misinterpreted.

I did refer the authors to the paper by Russell et al that explains why simply using a
standard error does not reflect the potential variability in deltaR values within a system.
They have misunderstood my meaning and simply explained how they calculated the
standard error. I suggest again that they use the standard error for predicted values as
the measure of the uncertainty in the deltaR values.

The ΔR error was recalculated in order to represent an approach more realistic of the
variability inherent in ΔR, according to the methodology proposed by Russell et al. 2011,
using the equation for the standard error for predicted values:

Standard error = Δ

This was explained in the document (line 389-392, page 20), the values required for the
calculation are shown in the table 2 and are corrected in Figure 7 and 8.
*Manuscript
Click here to view linked References

1 CALCULATION OF THE RESERVOIR AGE FROM ORGANIC AND CARBONATE

2 FRACTIONS OF SEDIMENTS IN THE GULF OF CARIACO (CARIBBEAN SEA)

4 Iliana Aguilara,b , Pierre Sabatierc, Christian Becka, Franck Audemardb, Christian Crouzeta,

5 Franco Urbanid, Corina Campose

7 iliana.aguilar-ramos@univ-savoie.fr

8 (a) CNRS ISTerre, Université Savoie Mont Blanc, 73376 Le Bourget du Lac, France.

9 (b) Venezuelan Foundation for Seismological Research, El Llanito, Caracas, Venezuela.

10 (c) CNRS EDYTEM, Université Savoie Mont Blanc, 73376 Le Bourget du Lac, France.

11 (d) Departamento de Geología, Universidad Central de Venezuela, Los Chaguaramos,

12 Caracas, Venezuela.

13 (e) Departamento de Ciencias de la Tierra, Universidad Simón Bolívar, Sartenejas, Caracas,

14 Venezuela.

15

14
16 ABSTRACT. A set of 24 AMS C measurements were performed on different sediment

17 fractions from short gravity cores taken the Gulf of Cariaco (Northeastern Venezuela), in

18 order to construct a detailed sedimentary archive for the last centuries. There, a local reservoir

19 effect was expected because of: i) strong upwelling from the neighbouring Cariaco Trough

20 and anoxia, ii) specific geomorphological setting related to active fault with seepages and

21 diapirism. Measurements was thus performed on different components: sediment bulk organic

22 fraction (SBOF), pteropods shells (surface water), bivalve shells (bottom), and plant

23 fragments. Based on comparisons of the different results, we discuss different age corrections

24 and calibrations, which lead us to consider a localized reservoir correction (ΔR) negligible for

25 bivalve and pteropod shells, showing no reservoir age for carbonate and no water

1
26 stratification. At the difference, a local 633 ± 64 yr ΔR appears recorded in the organic

27 fraction of sedimentation and its origin is primary related to the upwelling mechanism.

28

14
29 Keywords: late Holocene, Reservoir effect, C dating AMS, Gulf of Cariaco, LOI, ROCK-

30 EVAL Pyrolysis.

31

32 1. INTRODUCTION

33

34 The difference between radiocarbon ages of organisms belonging to the terrestrial biosphere

35 and their contemporary marine equivalents was defined by Stuiver and Polach (1977) as

36 “reservoir age". For aquatic (marine, lacustrine or mixed) organisms, if considering organic

37 matter (OM) coming from photosynthesis, the amount of fixed 14C depends, on one hand, on
14
38 its atmospheric production rate (changing with cosmic flux), and, on the other hand, on C
14
39 activity within incorporated dissolved CO2 in the seawater. At its turns, this C content

40 depends on CO2 residence time within the emphasized water mass, and to hard water effect
14
41 related C-free inorganic carbon originating from sub-aerial dissolution of old carbonate

42 rocks (Spennemann and Head, 1998; Sabatier et al., 2010). The delay between dissolution and

43 incorporation into marine OM is assigned to local different influences: water depth,

44 circulation, surface mixing, etc.

45 When the age difference between a marine sample and its atmospheric “stratigraphic”
14
46 equivalent is linked to marine C variations in response to CO2 exchanges between the

47 atmosphere and the ocean, this effect may be corrected subtracting a variable R t derived from
14
48 modelled ocean average (Reimer et al., 2013) from the conventional marine C age.

49 However, several studies have suggested the possibility of significant deviations in regional

50 marine reservoir signature from this average value (Folk and Ward, 1957; Goodfriend and

51 Flessa, 1997; Siani et al., 2001; Reimer and McCormac, 2002; Southon et al., 2002; Sabatier

2
52 et al., 2010). This ΔR value represents the exchange conditions of radiocarbon between the

53 atmosphere and a specific reservoir, and can be quantified by the difference between the

54 regional marine and the global marine 14C age (Stuiver and Braziunas, 1993).

55 For conventional ages derived from particulate (bulk) organic fraction from gulfs and

56 estuaries sediments, the corrections of overestimated ages due to this effect is not evident, and

57 can be linked to: i) possible mixing of reworked fluvial and marine organic matter (Ingram

58 and Southon, 1996; Goodfriend and Flessa, 1997), ii) hard water effect (Spennemann and

59 Head, 1998), iii) upwelling process (Stuiver and Braziunas, 1993), iv) carbon sequestration in

60 sediments and carbonate from the methanogenesis (Aharon and Gupta, 1994; Luff and

61 Wallmann, 2003; Orphan et al., 2004; Naehr et al., 2007; Paull et al., 2007).

62

63 In the case of the Gulf of Cariaco, no local reservoir effect had been previously reported,

64 neither in sediment bulk organic fraction (SBOF) nor in carbonate shells. We here present a

65 set of 24 radiocarbon dates taken from two cores collected from the deepest part of the gulf,

66 under anoxic conditions. We aimed to establish an accurate chronology for the last centuries,

67 in order to detect traces of historical earthquakes and tsunami (Aguilar et al., 2015).

68

69 2. STUDY SITE

70

71 The Gulf of Cariaco - an appendix of the Cariaco Basin (or Cariaco Trough) - is an E-W

72 elongated semi-closed basin, 60 km long and 15 km wide, with 55 m average depth; it is

73 located in Northeastern Venezuela, on the Southeastern boundary of the Caribbean realm

74 (Figure 1).

3
75

76 Figure 1. Geomorphological setting and location of short gravity cores. Main active trace of the El Pilar fault

77 following Audemard et al. (2000) and Audemard et al. (2007), simplified bathymetry from Caraballo (1982a).

78 Lithology and shaded-relief map of the watershed modified from Hackley et al. (2005). Numbers in white circles

79 indicate the main Southern drainage. Insert: geodynamic setting of the Cariaco Trough.

80

81 3. GEOLOGICAL SETTING

82

83 The Cariaco Basin is a pull apart which relays the San Sebastian Fault (West) and the El Pilar

84 Fault (East). The Gulf of Cariaco is directly built upon the later one (Figure 1). More

85 precisely, the El Pilar dextral strike slip fault (EPF) is represented by the VE-13b segment,

86 whose submarine morphologic expression has about 50 km through the Gulf (Audemard et

87 al., 2007). The bottom of the gulf is characterized by two singularities (Figure 1):

88 - along the south coast: the Guaracayal depression (GD). It is associated to the

89 submarine transtensive relay of the active trace of the El Pilar fault and it is defined as a pull-

4
90 apart basin by Audemard et al. (2007). It has a 90 m maximum depth and represents about

91 3.5% of the total gulf volume (Okuda, 1981);

92 - to the west, near the gulf mouth, the Salazar Sill (SS) represents the connection path

93 between the gulf and the deep Cariaco Basin (Caribbean Sea). The link between the Salazar

94 Sill (isobaths -50 to -40 m) and the Cariaco Basin is represented by a NE-SW trending

95 submarine canyon in the gulf (Caraballo, 1982b).

96 The lithological framework of the Gulf of Cariaco’s catchment area is divided into two

97 contrasting areas: to the north, the Araya Range formed mainly by Paleozoic to Cretaceous

98 metamorphic rocks; to the south, the sedimentary rocks of the Eastern Interior Range.

99 The Araya Peninsula separates the gulf from the Caribbean Sea, at its turn, it can be divided

100 into three zones: 1) its western tip with dominant recent alluvial and costal deposits:

101 limestones, marls and marines conglomerates, sandstones and shales, these lithologies

102 representing 10% the watershed; 2) the central zone formed principally by metamorphic

103 rocks: quartz-micaceous schists, quartz-chlorite phyllites and quartz-micaceous-calcareous

104 rocks (where the carbonate fraction represents 20% of the unit); 3) the eastern zone dominated

105 by quartz-micaceous-chloritic schists and sandy shales, shales and conglomeratic sands.

106 On the southern side, the Eastern Interior Range is represented by quartzitic sands and shales,

107 with minor limestones lenses.

108

109 4. HYDROGRAPHIC SETTING

110

111 In addition to the above mentioned lithological contrasts, there is also a clear difference in the

112 hydrographic network. Along the northern coast the Gulf’s tributaries are short with

113 intermittent flooding (Febres-Ortega, 1974; Caraballo, 1982b; Márquez et al., 2005; Quintero

114 et al., 2005). Due to low precipitations on the Araya Peninsula, theses rivers have a minor

5
115 influence within the whole hydrological input into the Gulf of Cariaco (Márquez et al., 2011).

116 On the southern coast, six short tributaries are considered as principal sources of runoff

117 waters, from west to east: the Tunantal (1), Guaracayal (2), Marigüitar (3), Tarabacoa (4),

118 Cachamaure (5) and Cariaco (6) rivers (Márquez et al., 2005). This coast shows a scarped

119 morphology controlled by arid conditions (Figure 1), where the evaporation exceeds largely

120 the precipitation rate (Quintero et al., 2005).

121 The hydrographic conditions in the Gulf of Cariaco are strongly influenced by the process of

122 water circulation and exchange with the Caribbean Sea. Waters flowing into the gulf from a

123 sub-surface layer of the Cariaco Basin charged of organic detritus have been observed (Okuda

124 and Benitez-Alvarez, 1974; Okuda et al., 1974), so part of the particulate organic matter here

125 accumulated come from the Cariaco basin.

126 The cyclical pattern of circulation into the Gulf of Cariaco is controlled by the trade-winds

127 and a thermal stratification of isohaline waters. The seasonality is follows the Inter-Tropical

128 Convergence Zone (ITCZ) migration (Figure 2A) along northern Venezuela. Thus, two

129 markedly different regimes characterize the hydrodynamic of oceanic water coming to the

130 gulf: 1) a weak upwelling period between July and December, where weak winds regime is

131 not favourable for the water mass movement (Figure 2B above) and 2) a strong upwelling

132 period during the rest of the year related to stronger trade-wind intensities that drive the

133 oceanic circulation (Figure 2B bottom).

134 The weak upwelling period (Figure 2B, above) named “stagnation” by Okuda et al. (1978) is

135 characterized by the stabilization of the water body under 50 m, while a horizontal mixing

136 process dominates in superficial water. This gulf mass water stratification into two units is

137 induced by the development of a distinct thermocline between 40-50 m depth, acting as a

138 barrier for the vertical water exchange process (Richards, 1960). The water at depths greater

139 than 60 m remains isolated, which leads to the formation of anoxic conditions.

6
140

141 Figure 2. Seasonal variation of upwelling related to the ITCZ. A) Location of the Gulf of Cariaco relative to the

142 ITCZ and the trade-winds position adapted from Poore et al. (2004) (Global relief model from Amante and

143 Eakins, 2009; Goni et al., 2009), B) Dynamic of the subsurface currents and the thermal balance into the gulf

144 during the period of strong and weak upwelling (following Gade, 1961; Okuda, 1981; Okuda, 1982); C) Relative

145 position of the cross section shown in B.

146

147 During periods of strong upwelling or renewal (Figure 2B, bottom), the thermocline is

148 relatively weak; sub-surface nutrient-rich waters from the Cariaco Basin penetrate into the

149 gulf displacing the coldest and densest bottom waters towards the east, where they emerge

150 increasing the outflow of superficial waters and thus the nutrients availability in the first few

151 meters of the water column (Okuda, 1982). The vertical range of this upwelling process was

152 limited by Griffiths and Simpson (1967) to the first 50 m and by Caraballo (1982b) between

153 the 30-60 m, both agree about the isolation of the bottom and the existence of stagnant water

154 below the sill depth a main part of the year. Richards (1960) does not discard the possibility

155 of exceptional renewal of bottom waters.

156 The superficial waters (photic zone) enriched in nutrients during the renewal period stimulate

157 biosynthesis processes and are responsible for the cyclic process associated to the high

158 primary productivity (231 g/m2.year according to Romero et al., 2013). Bonilla and Lin

7
159 (1979) suggest that the high concentration of organic matter in sediments of the gulf mainly

160 results from high primary productivity linked to the upwelling.

161 Century-scale analysis of the late Holocene sedimentary record from Cariaco Basin indicates

162 a major climatic change around the end of the Little Ice Age interpreted as a relative reduction

163 in local upwelling (Black et al., 1999; Black et al., 2004; Peterson and Haug, 2006; Polissar et

164 al., 2006).

165

166 5. METHODOLOGY

167

168 5.1. Sampling and sedimentary analyses

169

170 Two short gravity cores (around one meter) were selected among a set of 18 ones; they were

171 achieved in the Guaracayal depression: Cariac-09-06 (10°28’29”N; 63°57’58”W) and Cariac-

172 09-05b (10°18’19’’N; 63°58’57”W) in 2009, at a 90 meter water depth (Figure 1), on board

173 the R/V GUAÏQUERI II. At the laboratory, the cores were cut into two halves. Each half-

174 section was described in detail and pictures were taken. The lithological description of the

175 sequence allowed the identification of different sedimentary facies. The visual description of

176 smear slides of the sediment permitted better characterizing the different units present.

177 The Cariac-09-06 was sampled every 5 mm, and the Cariac-09-05B every 2.5 cm following

178 the methodology proposed by Heiri et al. (2001). The Loss On Ignition (LOI) analyses were

179 used to estimate the variations of organic and carbonate fraction contents in sediments,

180 performed in the oven Nabertherm Controller B180 of the Dynamic Environments and

181 Territory of the Mountain laboratory (EDYTEM). The weight loss during a first heating step

182 to 550°C corresponds to LOI550, while the second ignition phase to 950°C allows the LOI950

183 calculation. These values are respectively proportional to the amount of OM and carbonate.

8
184 The high resolution particle size analysis were performed in the Earth Science Intitute

185 laboratory (ISterre) with a laser diffraction microgranulometer MALVERN™ Mastersizer

186 2000 at a continuous interval of 1 cm for the whole cores and each 2 to 5 mm for selected

187 sections. Two statistic parameters were used (Mean size and Q99) to characterize the evolution

188 of grain size along cores.

189 In order to characterize the (OM) a set of 57 ROCK-EVAL pyrolysis measurements from

190 bulk sediment samples were carried every 2 cm from Cariac-09-06. Once dried to 60°C and

191 pulverized, the measurements of hydrogen index (HI), oxygen index (OI) and total organic

192 carbon (TOC) parameters were performed using a ROCK-EVAL 6 analyzer (RE6, Vinci

193 Technologies®) in the ISTO laboratory at the Orléans University following the method

194 described by Espitalie et al. (1985) and Lafargue et al. (1998). These values were used to

195 discriminate the source of OM in the bulk samples through the use of the modified Van

196 Krevelen diagram (Lafargue et al., 1998).

197

198 5.2. 14C dating

199

200 In the Gulf of Cariaco, three different types of material were sampled for dating purpose: 1)

201 Carbonate macro-rests: these samples correspond to well preserved pteropods (planktonic)

202 and bivalves shells (benthic); 2) Terrestrial macro-rests: only two samples were found and

203 collected, a leaf into the Cariac-09-06 and a well-preserved wood fragment in the Cariac-09-

204 10 core; 3) SBOF for each contemporary macro-remain (carbonated and terrestrial) was also

205 collected. These samples principally correspond to a silt-sized terrigenous fraction rich in

206 organic particulate matter, in which the carbonate content was removed by acid wash (HCl)
14
207 prior analysis. The 24 conventional C ages determined in this study were corrected by

208 isotopic fractionation, using the delta 13C values and processed at two laboratories, the Saclay

9
209 CEA Laboratory of the CNRS-INSU ARTEMIS program and in BETA ANALYTIC

210 laboratory.

211

212 6. RESULTS

213

214 6.1. Main sedimentation features

215

216 The two cores display a uniform olive green silty sediment, darker than most detrital sand

217 layers. These observations associated to grain size, LOI analysis and smear slide description

218 allow defining two facies:

219 -Facies 1: it is present in all cores. Its color is predominantly olive gray (5 Y 3/2) and is rich

220 in organic matter (≃19% wt). Particle size distributions (Figure 3) shows a dominance of silt-

221 sized particles between 15-30µm (Cariac-09-06: 70.0 ± 4.5 %; Cariac-09-05B: 63 ± 6 %),

222 lesser amounts of clay size (Cariac-09-06: 14 ± 3 %; Cariac-09-05B: 10.0 ± 1.6 %) and sand

223 size (Cariac-09-06: 16 ± 5 %; Cariac-09-05B: 27 ± 6 %); sorting is poor (3>So>2). Smear

224 slides display that silt-sized particles are a mixture of a lithic fraction (~70%) and different

225 kinds of biogenic materials (30%). The principal mineral grains are of: gypsum, quartz,

226 calcite, halite and mica. The biogenic fraction includes foraminifera (<2%), mollusc shell

227 fragments (<3%), diatoms (5%) and organic matter particle (20%). These characteristics

228 define the permanent (“background”) sedimentation, representing about 90% of the sediments

229 in studied cores.

230 - Facies 2: it is composed of olive gray silty sediment (5Y4/1), slightly lighter than facies 1,

231 with thicknesses between 2 and 4 cm, accompanied by macroscopic fibrous OM and leaf

232 fragments. This more terrigenous siliciclastic facies was identified three times in the core

233 collected from the deepest area of the gulf (Cariac-09-06). The most superficial level of the

10
234 three (40-45 cm deep) does not have a distinctive imprint on the mean grain size profiles (this

235 level slightly exceeds the average grain size of the whole core: 32 µm). However, it is

236 possible to distinguish this facies on the Q99 profile through a significant increase in grain

237 size of the coarse fraction (> 250 µm).

238 The Cariac-09-05B core, located 2 km west of the Cariac-09-06 core in the Guaracayal basin,

239 consists entirely of facies 1 (Figure 3B).

240

241 6.2. Loss on ignition (LOI)

242

243 Two profiles of LOI550, LOI950 and refractory residues vs. depth (Figure 3) illustrate the

244 variations of OM content, carbonate content and siliciclastic residues. The average amount of

245 OM in Cariac-09-06 corresponds to 19 % wt with a standard deviation of 0.5. This value does

246 not change abruptly along core, except for the facies 2 with an increase of siliciclastics

247 percentage between 70-80% and a consequent relative decrease in OM (12-16%).

11
248

249 Figure 3. Evolution of different particle size fractions (sand, silt, and clay), Q99, mean grain size and parameters

250 derived from the loss of ignition (LOI). Estimated weight percent of siliciclastic residues, carbonate (LOI 950) and

251 OM (LOI550) from Cariac-09-06 (a) and Cariac-09-05B (b) cores.

252

253 On the Cariac-09-05B core, the average amount of organic fraction is estimated at 19.7 % and

254 σ = 1.3 (from LOI550), very close to those values estimated from visual inspection of smear

255 slides. At 42 cm depth we observe an increase of the refractory fraction percentage. However,

256 this change is not eye-visible on the split core. In this core the carbonate percent remains

257 almost constant (6% average, σ = 1), in the same way as in the other core, the amount of

258 carbonate co-varies with OM.

259

260 6.3. Radiocarbon ages

261

12
14
262 Conventional C measurements performed on Cariac-09-06 and Cariac-09-05B cores are

263 presented in Table 1. Preliminarily the conventional ages obtained from carbonate shells and

264 SBOF were calibrated with R=0. Ages derived from shells were calibrated using the

265 Marine13 calibration curve.

266 The dated SBOF samples appear largely of marine origin. Following interpretations of

267 Espitalie et al. (1985), the different domains into Pseudo - Van Krevelen diagram (Hydrogen

268 Index (HI) vs. Oxygen Index (OI)) lead to the distinction of three types of organic matter:

269 Type I of lacustrine origin, type II which is of marine origin and type III, which is mostly

270 continental in origin. Langford and Blanc-Valleron (1990) propose a distribution of type II

271 organic matter unextractable on this diagram, constrained between 500 and 200 HI.

272 Consequently, considering the test results of ROCK EVAL pyrolysis (Figure 4) these samples

273 were calibrated also with the Marine13 curve.

274 The conventional age of the Cariac-09-06-1 sample (Table 1) is reported in percentage of

275 modern carbon (112.6 ± 0.3 pCM). Its negative equivalent radiocarbon age BP is calculated

276 with the R-code package to perform classic age-depth modelling Clam 2.2 (Blaauw, 2010)

277 and calibrated with the post-bomb curve corresponding to the zone 2 in the north hemisphere

278 (Hua et al., 2013). The atmospheric sample (Cariac-09-06_5 sample) was calibrated with the

279 curve INTcal13. The conventional ages from SBOF show inversions between 80.0 and 83.5

280 cm and 95 and 101 cm.

13
Figure 5. Plot of conventional ages for Cariac-09-06
Figure 4. Modified Van Krevelen diagram presenting
core (non-calibrated age). Ages in red correspond to
the results of ROCK-EVAL pyrolysis. It includes 57
marine carbonated samples, while in blue to organic
samples from Cariac-09-06 core. The samples located
fraction of the sediment. The sample in green
near the limit of Type III OM (red points) were
corresponds to the age of a terrestrial macro-rest.
collected from facies 2 at 84 and 95 cm depth.
ANE: Atmospheric Nuclear Experiments.

281

282 Table 1. Radiocarbon ages obtained from Cariac 09-06 and Cariac 09-05B cores

Conventional δ13C Marine13 Age cal. BP2


N° Sample name Lab. Depth Type1 Median Median
age (BP) o/oo (Cal. BP)2 (ΔR = 633 ± 64yr)

1 Cariac-09-06_1 BA 15.5 Ptp* -910 ± 21 N/A -8 to -47 -44 N/A N/A

2 Cariac-09-06_2 BA 15.5 SBOF 510 ± 30 -20.4 246-Post BP 0 135 N/A N/A

3 Cariac-09-06_3 BA 28.5 BS 400 ± 30 +0.8 103-Post BP 0 40 N/A N/A

4 Cariac-09-06_4 BA 28.5 SBOF 1080 ± 30 -20.2 695-565 646 232 - Post BP 0 92

5 Cariac-09-06_17 BA 34 BS 480 ± 30 N/A 226-Post BP 0 87 N/A N/A

6 Cariac-09-06a CEA-S 39 SBOF 1070 ± 30 -21 683-558 638 231 - Post BP 0 86

7 Cariac-09-06_5 BA 46 leaf** 150 ± 30 -28.4 283-Post BP 0 161 N/A N/A

8 Cariac-09-06_6 BA 46 SBOF 1230 ± 30 -20.4 868-691 768 403-54 224

9 Cariac-09-06_7 BA 52 Ptp 610 ± 30 N/A 309-141 258 N/A N/A

10 Cariac-09-06_8 BA 52 SBOF 1210 ± 30 -19.7 834-671 756 356-Post BP 0 199

11 Cariac-09-06_9 BA 65.5 BS 670 ± 30 NA 405-262 316 N/A N/A

14
12 Cariac-09-06_10 BA 65.5 SBOF 1210 ± 30 -19.7 834-671 737 336-Post BP 0 199

13 Cariac-09-06b CEA-S 80 SBOF 1100 ± 30 -20.3 723-609 659 N/A N/A

14 Cariac-09-06_11 BA 83.5 SBOF 1620 ± 30 -21.0 1263-1101 1191 N/A N/A

15 Cariac-09-06c CEA-S 94 SBOF 1490 ± 30 -20.3 1136-953 1036 N/A N/A

16 Cariac-09-06_12 BA 95 SBOF 1950 ± 30 -22.3 1586-1403 1502 N/A N/A

17 Cariac-09-06_13 BA 101 BS 1000 ± 30 +0.2 640-527 583 N/A N/A

18 Cariac-09-06_14 BA 101 SBOF 1670 ± 30 -19.2 1288-1168 1234 716-505 607

19 Cariac-09-06_15 BA 107 Ptp 1100 ± 30 N/A 723-609 658 N/A N/A

20 Cariac-09-06_16 BA 107 SBOF 1760 ± 30 -20.4 1373-1253 1306 826-547 685

21 Cariac0905B_30 BA 30 SBOF 1040 ± 30 -19.9 661-546 613 226 - Post BP 0 72

22 Cariac09_05B_61 CEA-S 61 SBOF 1390 ± 30 -20.1 1032-875 936 501-276 391

23 Cariac09_05B_91 CEA-S 91 SBOF 1700 ± 30 -19.7 1312-1182 1259 749-518 631

24 Cariac-09-05B CEA-S 95 SBOF 1820 ± 30 -20.1 1460-1285 1356 891-633 739

1
Abbreviations

BA: Beta Analytic; CEA-S: Saclay CEA Laboratory of the CNRS-INSU ARTEMIS; SBOF:

sediment bulk organic fraction; Ptp: Pteropod.


2
We used a range calibration of 95%
*
Age reported in percentage of modern carbon
**
Sample calibrated with the curve INTcal13

283 .

284 The depth distribution of uncalibrated ages (Figure 5) shows a linear pattern controlled by the

285 nature of the dated material. Ages in red are carbonate samples: bivalves shell (benthic) and

286 pteropod (planktonic), the SBOF samples are indicated in blue and the only atmospheric
14
287 sample in green. Both shells and SBOF sediment C ages, mostly follow a

288 chronostratigraphic order, but the relatively older dates of bulk samples suggest a reservoir

289 age effect.

290 The offset with the zero of the model is approximately 390 yr for carbonated samples and

291 1000 yr for SBOF samples. The grey bars, represent the facies 2 deposits. The horizontal

292 arrow indicates the quasi-constant shift between the ages of SBOF and its marine coeval

293 sample, this offset is equivalent to ΔR.

15
294

295 7. DISCUSSIONS

296

297 7.1. Radiocarbon anomalies

298

299 The age of the most superficial samples (above 15 cm depth in core) will not be considered in

300 this calculation. The modern organic carbon content in the sample of marine origin collected

301 within the first 15 cm implies that the sediments placed to this same depth may potentially

302 record an increase in the 14C activity associated to post-bomb peak. Consequently, the global

303 and local reservoir effect correction would be wrong.

304 Facies 2 type sediments (grey bands in Figure 5) present an increase of siliciclastic material

305 (Figure 3a). This facies could be correlated with two coarse siliciclastic layers (fine sand)

306 identified in a core close to the south coast (Cariac-09-18; see Figure 1 for location). In these

307 layers (in the proximal zone) the base-to-top evolution of grain size have the structure of a

308 hyperpycnal flow composed of an upward-coarsening basal unit, an upward-fining top unit

309 and an internal maximal grain size in between. Thus the facies 2 is considered as the distal

310 facies of a terrigenous input interpreted as flood event deposits (Aguilar et al., 2015).

311 In these deposits, dated SBOF samples present an age inversion and thus have to be removed

312 from the ΔR calculation. The significant offsets from the general trend recorded between 82-

313 84 cm and 94-98 cm depths in the age model are considered as a consequence of dating of old

314 reworked organic matter material of terrestrial origin (Figure 4) transported during settling of

315 flood deposits.

316

317 7.2. Modern marine ΔR age estimation on carbonated sample

318

16
319 In the Gulf of Cariaco the local reservoir effect had not been reported previously, either for

320 organic sediments or for carbonates. From dating of monospecific samples of planktonic

321 foraminifera collected inside the sub-basin east of the Cariaco Basin (900 m depth, Figure 1),

322 Hughen et al. (2004) proposed R(t) = 430 ± 30 and ΔR = 20 values. Compared with open

323 North Atlantic values (435 ± 31) (Mangerud et al., 2006), this age reservoir shows no

324 significant offset. This fit between 14C marine reservoir age recorded in planktonic samples of

325 the Cariaco basin and the open sea values has been explained by well-equilibrated dissolved

326 CO2 in surface waters with respect to the atmosphere (Hughen et al., 1998).

327

328 A minor difference among the global mean sea surface reservoir ages and the values yielded

329 by foraminifera in the deep Cariaco Basin, leads us to suppose that shells samples from the

330 Gulf of Cariaco could record negligible ΔR values. This assumption is supported by the fact

331 that water exchange between open sea and the gulf concerns the sub-surface water (<120 m;

332 Richards, 1960). For this reason, the Marine13 curve (Reimer et al., 2013) was used here to

333 calibrate carbonated samples (pteropods and bivalves) from Cariac-09-06 core, without

334 considering any local correction (ΔR= 0).

335

336 The age-depth curve for the Cariac-09-06 core, built with ages from marine carbonated

337 samples (Figure 7) both planktonic and benthonic organisms, does not present significant age

338 shift, suggesting that the superficial and deep waters in the gulf do not record an apparent ΔR.

339 The absence of a positive offset, between ages obtained at the top of core and the zero of the

340 model - after the correction with the Marine13 curve - leads us to suppose that the ΔR

341 recorded in carbonates is negligible. In addition to this argument, the fit between of the only

342 “atmospheric” sample recovered within the gulf, with the age vs. depth curve, allows us to

343 presume that there is not ΔR value to suppress. This fact must be associated to the reduced

17
344 depth in the gulf (<100m) and to exchange process of superficial and sub-superficial waters

345 with Caribbean Sea well-equilibrated with the atmosphere.

346

347 7.3. Modern marine ΔR age estimation on particulate organic matter sample

348

349 Plot of Figure 4 underlines that all ages obtained from SBOF samples are yielding an

350 overestimation. But this age shift is relatively constant along the core with respect to its

351 coeval marine carbonate sample. This result leads us to propose the existence of a constant

352 ΔR value to correct all SBOF samples.

353 The regional reservoir age is defined by the expression R'(t) = global R(t) + ΔR, where R(t)

354 term is incorporated into the global marine curve and ΔR is equivalent to the regional
14
355 reservoir age offset from ocean C age. This ΔR value can be calculated based on pairs of

356 coeval samples, following Stuiver and Braziunas (1993):

357 ΔR = conventional age of marine sample - Marine 14C ages calculated from atmospheric ages (1)

358 In order to calculate the ΔR recorded by the SBOF samples, we used: 1) the conventional

359 radiocarbon age from the plant fragment (A = 150 ± 30); 2) the equivalent marine radiocarbon

360 age derived from the atmospheric sample (B = 536 ± 80); 3) the measured SBOF

361 conventional radiocarbon age (C = 1230 ± 30 yr BP).


14
362 These values are shown in the C age vs. Cal. ages diagram of Figure 6, where the curves

363 Marine13 and Intcal13 are plotted. Here, the atmospheric (leaf) sample is first calibrated with

364 the atmospheric calibration curve (Reimer et al., 2013), its cal. age is then projected vertically

365 upon the Marine 13 curve (Reimer et al., 2013) and converted to a equivalent marine

366 radiocarbon age reprojected on the vertical axis of the diagram.

367 The ΔR value is the difference between the SBOF conventional radiocarbon age and the

368 marine radiocarbon age equivalent to the sample of atmospheric origin (Figure 6):

369 ΔR = C – B (2)

18
370
371 Figure 6. Conversion of 14C conventional ages into calendar ages for sub-aerial samples equilibrated with

372 atmosphere (lower curve) and marine samples (upper curve). The age difference between the 2 curves for a

373 given calendar date corresponds to the marine reservoir age R(t). Dating of the terrestrial macro-remains at 150 ±

374 30 (A) gives a marine conventional 14C age of 536 ± 80 yr and thus a reservoir age of 1230−150 = 1080 yr.

375 Difference from the global mean reservoir age (ΔR) for SBOF sample is obtained by subtracting the marine

376 model age value estimated at the calendar date from the measured apparent 14C age, thus 1230−536 = 694 yr.

377

378 Assuming that ΔR value recorded in the shells is null, the B term of the formula (2) for the

379 ΔR estimation can be substituted by conventional marine ages derived from dated marine

380 carbonate samples. Thus, we proceed to estimate the average value of the local reservoir

381 effect registered into the SBOF using the following equation:

382 ΔR = SBOF conventional age - conventional marine age from coeval marine carbonate sample (3)

383 Six pairs of samples provide a weighted mean ΔR = 633 ± 24 (with an error on the weighted

384 mean, n = 6, ± 1 σ) calculated following Scott et al. (2007), which are presented in Table 2.

385

386

387 Table 2. ΔR calculation.

Sample Depth Type Conventional ΔR (yr)

19
(cm) age14C BP

Cariac-09-06_3 (388630) 28.5 Bivalve shell 400 ± 30 680 ± 42


Cariac-09-06_4 (388631) SBOF 1080 ± 30
Cariac-09-06_5 (388632) leaf* 536 ± 80
46 694 ± 85
Cariac-09-06_6 (388633) SBOF 1230 ± 30
Cariac-09-06_7 (388634) Pteropod shell 610 ± 30
52 600 ± 42
Cariac-09-06_8 (388635) SBOF 1210 ± 30
Cariac-09-06_9 (388636) SBOF 670 ± 30
65.5 540 ± 42
Cariac-09-06_10(388637) SBOF 1210 ± 30
Cariac-09-06_13(388640) Bivalve shell 1000 ± 30
101 670 ± 42
Cariac-09-06_14(388641) SBOF 1670 ± 30
Cariac-09-06_15(388642) Pteropod shell 1100 ± 30
107 660 ± 42
Cariac-09-06_16(388643) SBOF 1760 ± 30
Weighted mean ΔR 633 ± 24
StDev on the ΔR 59
*The reported age corresponds to its marine equivalent.

388

389 The error on the weighted mean obtained was considered too low (±24, table 2). Here we use

390 the standard error for predicted values (formula 4) proposed by Russell et al. (2011) in order

391 to represent an approach more realistic of the variability inherent in ΔR calculations:

392 Standard error = error on the weighted mean 2 standard deviation on the ΔR values 2 (4)

393 Using the standard error for predicted values the ΔR value is 633±64. In order to evaluate the

394 ΔR value, on Figure 7 we represented the age-depth curve from SBOF samples (blue) and the

395 age-depth curve from marine carbonate samples (black and grey) derived from Cariac-09-06

396 core, eliminating the inversed ages associated to flood events. The calibration of conventional

397 ages from carbonated samples (benthic and planktonic specimens) with Marine13 curve

398 seems to be enough to build a coherent age-depth curve. This curve shows a linear trend

399 without any abrupt changes in the sedimentation rate. Even the correction of the age obtained

400 from pteropods collected at 15 cm depth-in-core fits. The extrapolation of the age-depth curve

401 up to the top of the core, points out to an age close to the coring period (2009). We propose

402 that ΔR values in carbonates samples can be kept, because of: 1) the absence of a positive

20
403 offset between age at the top of core, 2) the alignment of the only atmospheric age recovered

404 within the gulf in the age model trend.

405

406 The corrections of reservoir effect (ΔR = 633 ± 64 yr) before calibration are thus applied to

407 the radiocarbon dates obtained from SBOF, and a constant sedimentation rate of 1.6 mm/yr is

408 deduced (Figure 7); it is close to the one derived from marine carbonates samples (1.7

409 mm/yr). The offset between zero of the model and the top of the core disappears. The

410 superposition of the two age-depth curves shows that the deviation between both are less than

411 the radiocarbon dates uncertainties (<30 yr). Therefore, we consider that the corrected ages

412 derived from SBOF can be used to perform the chronological sedimentation control within

413 the Gulf of Cariaco.

414 In a first approximation, the first 30 cm of the sedimentary record were discarded in the

415 calculation of ΔR under the suspicion that age from SBOF sample was affected by the modern
14
416 increased C activity linked to atmospheric weapon tests. The later ones leads to a decrease

417 of ΔR recorded in the OM as observed on the superficial level of the uncalibrated ages

418 (Figure 5). The here-studied sedimentary sequence seems to record a constant ΔR in the

419 period between 600-100 yr cal. BP.

21
420

421

422 Figure 7. Comparison between age-depth curves of marine carbonate samples (black line) and sediment

423 particulate organic fraction (SBOF) samples (blue line), performed with “clam R code package” from samples

424 collected of Cariac-09-06 core. Age of terrestrial macro-rest is in red.

425

426 7.4. Potential source of reservoir effect in sediment bulk organic fraction

427

428 Due to scarcity of carbonated lithological units in the watershed, a limited dissolution and

429 hydrolysis process in semi-arid environments and the absence of ΔR on marine carbonate, the

430 hard water effect is considered negligible on the estimated value of ΔR for SBOF.

431

432 On the other hand, gas emissions associated with faults and known hydrocarbon seepages

433 reported in the gulf along the southwestern coast of the Araya Peninsula, are considered as

434 potential old carbon sources. The methanotrophic bacteria in the organic matter could

435 incorporate, into its structure, 14C depleted methane recording anomalously old 14C age in this

436 type of samples (Reeburgh and Heggie, 1977; Fontugne et al., 2009).

22
437 An alternative explanation to the local reservoir effect recorded in the organic components of

438 the sediment deeper than 30 cm could be the strong trade wind-driven upwelling (Berger et

439 al., 1966; Taylor and Berger, 1967; Goodfriend and Flessa, 1997) reported in this Caribbean

440 zone during the Little Ice Age (Black et al., 1999; Polissar et al., 2006). The supply of

441 transported organic detritus transported (Okuda, 1981) into the gulf by upwelling from the

442 Cariaco Basin may be thus combined to the settling of local planktonic production.

443 Terrigenous siliciclastics levels have been identified and associated to continental sediments

444 input which may also contain reworked organic matter. Such a mixing with in situ OM may

445 lead to: either an underestimated age (terrestrial plants immediately reworked after growth,

446 with contemporaneous atmospheric CO2), or an overestimated ages (long time lapse between

447 growth and reworking).

448

449 The possible decrease of the reservoir effect in the most recent sediments (above 30 cm depth-

450 in-core) could be explained by an increase of terrestrial OM which is supposed to directly fix

451 the atmospheric CO2. Nevertheless, this explanation seems unlikely, based on the following

452 reasons. First, we found evidences for a mainly marine OM settling in the Gulf of Cariaco

453 (Aguilar et al., 2015). Second, according to different investigations (Black et al., 2004;

454 Peterson and Haug, 2006; Polissar et al., 2006), the end of the Little Ice Age is characterized,

455 in the studied area, by a decrease of upwelling. Thus, we favour a modification of seawater

456 exchange rather related to morphological changes (Aguilar et al., 2015).

457

458 7.5. Validation of ΔR estimation based on another core dating

459

460 In order to validate the proposed calibration parameters, the ΔR value calculated from Cariac-
14
461 09-06 SBOF samples was used to correct the C ages derived of organic samples collected

23
462 from another core, close to the previous one (Cariac-09-05B, Figure 1 for location). Figure 8

463 shows the evolution of age-depth relationships with the application of successive corrections

464 to non calibrated values. The age vs. depth curve on the right (dashed green line) corresponds

465 to conventional ages, the intermediate model (solid blue line) results from the application of

466 global reservoir correction with Marine13 calibration curve. The ages displayed by the red

467 curve have the correction of both R(t) an the above estimated ΔR. An average sedimentation

468 rate (1.6 mm/yr) very similar to the one estimated for the Cariac-09-06 core is obtained. The

469 absence of age offset at the top confirms the potential use of the local reservoir effect value

470 calculated in this study for radiocarbon age calibration in the Gulf of Cariaco.

471

472 Figure 8.Age-depth curve of Cariac-09-05B core. Dashed line: conventional 14C ages, blue dashed line:

473 calibrated 14C with Clam code package and Marine13 curve. Red dashed line: calibrated ages with the Marine13

474 curve and the ΔR = 633 ± 64 years.

475

476 8. CONCLUSION

477

24
478 The analyses presented in this work provide a ΔR estimation than can be used for correction

479 of 14C ages of SBOF samples from the Gulf of Cariaco. Our results indicate that, despite the
14
480 effect of a significant upwelling regime, C dating of different carbonated samples (benthic

481 and planktonic) do not record local reservoir effect (ΔR = 0). The average ΔR value derived

482 from sediment bulk particulate organic fraction 14C measurements was successfully applied to

483 produce a geochronology for sediment cores, and seem to be constant over the last 600 cal.

484 BP. Two explanations can be proposed to explain the high value of ΔR registered in the

485 organic fraction of the sediments: 1) the contribution of 14C-depleted organic detritus from the

486 upwelling, 2) the incorporation, within bottom microbial food chain, of possible 14C-depleted

487 methane seeping along the active El Pilar fault. Our analyses favour a major impact of

488 upwelling.

489

490 ACKNOWLEDGEMENTS

491

492 The Gulf of Cariaco coring survey performed in 2009 is part of a Venezuelan-French ECOS-

493 Norte grant N° V10U01, untitled "Contribution to multi-risk assessment - tsunamis,

494 earthquakes, extreme-weather events - along the edge Caribbean of Venezuela, through a

495 tectonic approach combined sedimentological and morphodynamic". We thank FUNVISIS,

496 CNRS-ISTerre, CNRS-BRGM-ISTO and CNRS-EDYTEM laboratories for cooperation and

497 laboratory facilities. Our investigations also benefit from CNRS-INSU ARTEMIS facility for
14
498 C dating. I. Aguilar’s stay in ISTerre Laboratory was funded through a FUNDAYACUCHO

499 PhD grant 756521F. Finally, many thanks to the whole crew of R/V GUAIQUERI II for their

500 unconditional and effective technical support and Andres Lemus, head of CAMUDOCA.

501

502 REFERENCES

25
503 Aguilar, I., Beck, C., Audemard, F., Develle, A., Boussafir, M., Campos, C., Crouzet, C., 2015. Last
504 millenium sedimentation in the Gulf of Cariaco (NE Venezuela): evidence for morphological changes
505 of Gulf entrance and possible relations with large earthquakes. CR Geoscience., in press.
506 Aharon, P., Gupta, B.K.S., 1994. Bathymetric reconstructions of the Miocene-age “calcari
507 aLucina”(northern Apennines, Italy) from oxygen isotopes and benthic Foraminifera. Geo-Marine
508 Letters 14, 219-230.
509 Amante, C., Eakins, B.W., 2009. ETOPO1 1 arc-minute global relief model: procedures, data sources
510 and analysis. US Department of Commerce, National Oceanic and Atmospheric Administration,
511 National Environmental Satellite, Data, and Information Service, National Geophysical Data Center,
512 Marine Geology and Geophysics Division.
513 Audemard, F., Beck, C., Moernaut, J., De Rycker, K., De Batist, M., Sánchez, J., González, M.,
514 Sánchez, C., Versteeg, W., Malavé, G., 2007. La depresión submarina de Guaracayal, estado Sucre,
515 Venezuela: una barrera para la propagación de la ruptura cosísmica a lo largo de la falla de El pilar.
516 Interciencia 32, 735-741.
517 Audemard, F.A., Machette, M.N., Cox, J.W., Dart, R.L., Haller, K.M., 2000. Map of Quaternary Faults
518 of Venezuela. Map and Database of Quaternary Faults in Venezuela and Offshore regions. USGS
519 Open-File Report, 78.
520 Berger, R., Taylor, R., Libby, W., 1966. Radiocarbon content of marine shells from the California and
521 Mexican west coast. Science 153, 864-866.
522 Blaauw, M., 2010. Methods and code for ‘classical’age-modelling of radiocarbon sequences.
523 Quaternary Geochronology 5, 512-518.
524 Black, D.E., Peterson, L.C., Overpeck, J.T., Kaplan, A., Evans, M.N., Kashgarian, M., 1999. Eight
525 centuries of North Atlantic Ocean atmosphere variability. Science 286, 1709-1713.
526 Black, D.E., Thunell, R.C., Kaplan, A., Peterson, L.C., Tappa, E.J., 2004. A 2000‐year record of
527 Caribbean and tropical North Atlantic hydrographic variability. Paleoceanography 19.
528 Bonilla, J., Lin, A., 1979. Materia orgánica en los sedimentos de los Golfos de Paria y Cariaco,
529 Venezuela. Bol. Inst. Oceanogr. Univ. Oriente 18, 37-52.
530 Caraballo, L.F., 1982a. El Golfo de Cariaco. I: Morfologia y batimetria submarina. Estructuras y
531 tectonismo reciente. Bol. Inst. Oceanogr. Venezuela 21, 13-35.
532 Caraballo, L.F., 1982b. El Golfo de Cariaco. II: Los sedimentos superficiales, y su distribución por el
533 fondo. Fuente de sedimentos. Análisis mineralógico. Bol. Inst. Oceanogr. Venezuela 21, 37-65.
534 Espitalie, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock-Eval et ses applications. Deuxième
535 partie. Oil & Gas Science and Technology 40, 755-784.
536 Febres-Ortega, G., 1974. Circulación de las aguas superiores de la Fosa de Cariaco en abril de 1974.
537 Bol. Inst. Oceanogr. Venezuela 13, 79-86.
538 Folk, R.L., Ward, W.C., 1957. Brazos River bar: a study in the significance of grain size parameters. J.
539 of Sediment. Res. 27.
540 Fontugne, M., Guichard, F., Bentaleb, I., Strechie, C., Lericolais, G., 2009. Variations in 14C reservoir
541 ages of Black Sea waters and sedimentary organic carbon during anoxic periods: influence of
542 photosynthetic versus chemoautotrophic production. Radiocarbon 51, 969-976.
543 Gade, H., 1961. Further hydrographic observation in the Gulf of Cariaco, Venezuela, The circulation
544 and water exchange. Bol. Inst. Oceanogr. Venezuela 1, 359-395.
545 Goni, M., Aceves, H., Benitez-Nelson, B., Tappa, E., Thunell, R., Black, D., Muller-Karger, F., Astor,
546 Y., Varela, R., 2009. Oceanographic and climatologic controls on the compositions and fluxes of
547 biogenic materials in the water column and sediments of the Cariaco Basin over the Late Holocene.
548 Deep-Sea Res Pt I 56, 614-640.
549 Goodfriend, G.A., Flessa, K.W., 1997. Radiocarbon reservoir ages in the Gulf of California: roles of
550 upwelling and flow from the Colorado River. Radiocarbon 39, 139-148.
551 Griffiths, R., Simpson, J., 1967. Temperature structure of the Gulf of Cariaco, Venezuela, from Aug.
552 1959 to Aug. 1961. Ministerio de Agricultura y Cria, Caracas. Ser. Recurs. Explot. pesq 1, 161-169.
553 Hackley, P.C., Urbani, F., Karlsen, A.W., Garrity, C.P., 2005. Geologic shaded relief map of
554 Venezuela. USGS Open-File Report.
555 Heiri, O., Lotter, A.F., Lemcke, G., 2001. Loss on ignition as a method for estimating organic and
556 carbonate content in sediments: reproducibility and comparability of results. J. of Paleolimnol. 25, 101-
557 110.
558 Hua, Q., Barbetti, M., Rakowski, A.Z., 2013. Atmospheric radiocarbon for the period 1950–2010.
559 Radiocarbon 55, 2059-2072.
560 Hughen, K.A., Overpeck, J.T., Lehman, S.J., Kashgarian, M., Southon, J., Peterson, L.C., Alley, R.,
561 Sigman, D.M., 1998. Deglacial changes in ocean circulation from an extended radiocarbon calibration.
562 Nature 391, 65-68.

26
563 Hughen, K.A., Southon, J.R., Bertrand, C.J., Frantz, B., Zermeno, P., 2004. Cariaco basin calibration
564 update: Revisions to calendar and 14C chronologies for core PL07-58PC. Radiocarbon 46, 1161-
565 1187.
566 Ingram, B.L., Southon, J.R., 1996. Reservoir ages in eastern Pacific coastal and estuarine waters.
567 Radiocarbon 38, 573-582.
568 Lafargue, E., Marquis, F., Pillot, D., 1998. Rock-Eval 6 applications in hydrocarbon exploration,
569 production, and soil contamination studies. Oil & Gas Sci. Technol. 53, 421-437.
570 Langford, F., Blanc-Valleron, M.-M., 1990. Interpreting Rock-Eval Pyrolysis Data Using Graphs of
571 Pyrolizable Hydrocarbons vs. Total Organic Carbon (1). AAPG Bulletin 74, 799-804.
572 Luff, R., Wallmann, K., 2003. Fluid flow, methane fluxes, carbonate precipitation and biogeochemical
573 turnover in gas hydrate-bearing sediments at Hydrate Ridge, Cascadia Margin: numerical modeling
574 and mass balances. Geochim. Cosmochim. Ac. 67, 3403-3421.
575 Mangerud, J., Bondevik, S., Gulliksen, S., Hufthammer, A.K., Høisæter, T., 2006. Marine 14 C
576 reservoir ages for 19th century whales and molluscs from the North Atlantic. Quaternary Science
577 Reviews 25, 3228-3245.
578 Márquez, A., Bonilla, J., Martínez, G., Senior, W., Aguilera, D., González, Á., 2005. Estudio
579 geoquímico de los sedimentos superficiales del litoral nororiental del Golfo de Cariaco, Estado Sucre,
580 Venezuela. Bol. Inst. Oceanogr. Venezuela 44, 89-103.
581 Márquez, A., Senior, W., Benítez, A., Fermín, I., Martínez, G., González, Á., Castañeda, J., Alcalá, L.,
582 De La Cruz, R., 2011. Sector oriental del golfo de Cariaco, Venezuela. Una descripción de su
583 dinámica hidroquímica, procesos, y del rol de la surgencia costera estacional. Bol. Inst. Oceanogr.
584 Venezuela 50, 255-272.
585 Naehr, T.H., Eichhubl, P., Orphan, V.J., Hovland, M., Paull, C.K., Ussler, W., Lorenson, T.D., Greene,
586 H.G., 2007. Authigenic carbonate formation at hydrocarbon seeps in continental margin sediments: a
587 comparative study. Deep Sea Research Part II: Topical Studies in Oceanography 54, 1268-1291.
588 Okuda, T., 1981. Water exchange and the balance of phosphate in the Gulf of Cariaco, Venezuela.
589 Coastal Upwelling, 274-281.
590 Okuda, T., 1982. Rate of water renewal and phosphate input in the gulf of Cariaco, Venezuela. Bol.
591 Inst. Oceanogr. Venezuela 21, 3-12.
592 Okuda, T., Benitez-Alvarez, J., 1974. Condiciones hidrográficas de las capas superiores en la Fosa de
593 Cariaco y áreas adyacentes durante la época lluviosa. Bol. Inst. Oceanogr. Venezuela 13, 147-162.
594 Okuda, T., Benitez-Alvarez, J., Bonilla, J., Cedeño, G., 1978. Características hidrográficas del golfo de
595 Cariaco, Venezuela. Bol. Inst. Oceanogr. Venezuela 17, 69-88.
596 Okuda, T., Ruiz, J.B., Garcia, A., 1974. Algunas caracteristicas bioquimicas en el agua de la Fosa de
597 Cariaco. Bol. Inst. Oceanogr. Venezuela 13, 163-174.
598 Orphan, V., Ussler, W., Naehr, T., House, C., Hinrichs, K.-U., Paull, C., 2004. Geological,
599 geochemical, and microbiological heterogeneity of the seafloor around methane vents in the Eel River
600 Basin, offshore California. Chemical Geology 205, 265-289.
601 Paull, C.K., Ussler, W., Peltzer, E.T., Brewer, P.G., Keaten, R., Mitts, P.J., Nealon, J.W., Greinert, J.,
602 Herguera, J.-C., Perez, M.E., 2007. Authigenic carbon entombed in methane-soaked sediments from
603 the northeastern transform margin of the Guaymas Basin, Gulf of California. Deep Sea Research Part
604 II: Topical Studies in Oceanography 54, 1240-1267.
605 Peterson, L.C., Haug, G.H., 2006. Variability in the mean latitude of the Atlantic Intertropical
606 Convergence Zone as recorded by riverine input of sediments to the Cariaco Basin (Venezuela).
607 Palaeogeogr. Palaeocl. 234, 97-113.
608 Polissar, P., Abbott, M., Wolfe, A., Bezada, M., Rull, V., Bradley, R.S., 2006. Solar modulation of Little
609 Ice Age climate in the tropical Andes. Proceedings of the National Academy of Sciences 103, 8937-
610 8942.
611 Poore, R., Quinn, T., Verardo, S., 2004. Century‐scale movement of the Atlantic Intertropical
612 Convergence Zone linked to solar variability. Geophysical Research Letters 31.
613 Quintero, A., Terejova, G., Bonilla, J., 2005. Morfología costera del Golfo de Cariaco, Venezuela. Bol.
614 Inst. Oceanogr. Venezuela 44, 133-142.
615 Reeburgh, W.S., Heggie, D.T., 1977. Microbial methane consumption reactions and their effect on
616 methane distributions in freshwater and marine environments1. Limnology and Oceanography 22, 1-9.
617 Reimer, P., McCormac, F., 2002. Marine radiocarbon reservoir corrections for the Mediterranean and
618 Aegean Seas. Radiocarbon 44, 159-166.
619 Reimer, P.J., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Ramsey, C.B., Buck, C.E., Cheng, H.,
620 Edwards, R.L., Friedrich, M., 2013. IntCal13 and Marine13 radiocarbon age calibration curves 0–
621 50,000 years cal BP. Radiocarbon 55, 1869-1887.

27
622 Richards, F.A., 1960. Some chemical and hydrographic observations along the north coast of South
623 America—I. Cabo Tres Puntas to Curacao, including the Cariaco Trench and the Gulf of Cariaco.
624 Deep Sea Research (1953) 7, 163-182.
625 Romero, D., Martínez, G., Brito, F., Rodríguez, E., 2013. Estudio de línea base en la determinación de
626 hidrocarburos aromáticos policíclicos totales en sedimentos superficiales del sector oriental del Golfo
627 de Cariaco, Venezuela. Avances en Química 8, 47-54.
628 Russell, N., Cook, G., Ascough, P., Scott, E., Dugmore, A., 2011. Examining the inherent variability in
629 ΔR: New methods of presenting ΔR values and implications for MRE studies. Radiocarbon 53, 277-
630 288.
631 Sabatier, P., Dezileau, L., Blanchemanche, P., Siant, G., Condomines, M., Bentaleb, I., Piquès, G.,
632 2010. Holocene variations of radiocarbon reservoir ages in a Mediterranean lagoonal system.
633 Radiocarbon 52, 91.
634 Scott, E.M., Cook, G.T., Naysmith, P., 2007. Error and uncertainty in radiocarbon measurements.
635 Radiocarbon 49, 427-440.
636 Siani, G., Paterne, M., Michel, E., Sulpizio, R., Sbrana, A., Arnold, M., Haddad, G., 2001.
637 Mediterranean Sea surface radiocarbon reservoir age changes since the last glacial maximum.
638 Science 294, 1917-1920.
639 Southon, J., Kashgarian, M., Fontugne, M., Metivier, B., Yim, W.W., 2002. Marine reservoir
640 corrections for the Indian Ocean and Southeast Asia. Radiocarbon 44, 167-180.
641 Spennemann, D.H., Head, M.J., 1998. Tongan pottery chronology, 14C dates and the hardwater
642 effect. Quaternary Science Reviews 17, 1047-1056.
643 Stuiver, M., Braziunas, T.F., 1993. Modeling Atmospheric 14C Influences and 14C Ages of Marine
644 Samples to 10,000 BC. Radiocarbon 35, 137-189.
645 Stuiver, M., Polach, H.A., 1977. Discussion; reporting of C-14 data. Radiocarbon 19, 355-363.
646 Taylor, R., Berger, R., 1967. Radiocarbon content of marine shells from the Pacific coasts of Central
647 and South America. Science 158, 1180-1182.

648

28
Figure_1 TIFF
Click here to download high resolution image
Figure_2 TIFF
Click here to download high resolution image
Figure_3 EPS
Figure_4 EPS
Figure_5 EPS
Figure_6 EPS
Figure_7 EPS

14
-8 to -47 cal C cal yr BP
20
14
232 to 0 cal 14C cal yr BP
103 to 0 cal C cal yr BP
14
226 to 0 cal C cal yr BP

40
14
403-54 cal 14C cal yr BP
Depth (cm)

283 to 0 cal C cal yr BP


14
336-0 cal C cal yr BP
14
309-141 cal C cal yr BP

60
14
405-262 cal C cal yr BP
14
336-0 cal C cal yr BP
Shell model age
Bulk organic fraction
model age
80 Pteropod
Bivalve shell
SBOF
Leaf calibrated age
Instantaneous events 14
100 716-505 cal 14C cal yr BP
Dated sample 640-527 cal C cal yr BP
14
826-547 cal 14C cal yr BP
723-609 cal C cal yr BP

0 500 1000
cal BP
Figure_8 EPS

0 500 1000 1500 2000


0
20

30 cm = 661-546 cal. yr BP

1040 ± 30 yr BP
30 cm =226-0 cal. yr BP
40
Depth (cm)

61 cm = 1032-875 cal. yr BP
60

1390 ± 30 yr BP
61 cm = 501-276 cal. yr BP
80

91 cm = 749-518 cal. yr BP 91 cm = 1312-1182 cal. yr BP


1700 ± 30 yr BP

95 cm = 891-633 cal. yr BP 95 cm = 1460-1285 cal. yr BP 1820 ± 30 yr BP

cal BP
Table_1

Table 1. Radiocarbon ages obtained from Cariac 09-06 and Cariac 09-05B cores
Conventional δ13C Marine13 Age cal. BP2
1
N° Sample name Lab. Depth Type Median Median
age (BP) o/oo (Cal. BP)2 (ΔR = 633 ± 64yr)

1 Cariac-09-06_1 BA 15.5 Ptp* -910 ± 21 N/A -8 to -47 -44 N/A N/A

2 Cariac-09-06_2 BA 15.5 SBOF 510 ± 30 -20.4 246-Post BP 0 135 N/A N/A

3 Cariac-09-06_3 BA 28.5 BS 400 ± 30 +0.8 103-Post BP 0 40 N/A N/A

4 Cariac-09-06_4 BA 28.5 SBOF 1080 ± 30 -20.2 695-565 646 232 - Post BP 0 92

5 Cariac-09-06_17 BA 34 BS 480 ± 30 N/A 226-Post BP 0 87 N/A N/A

6 Cariac-09-06a CEA-S 39 SBOF 1070 ± 30 -21 683-558 638 231 - Post BP 0 86

7 Cariac-09-06_5 BA 46 leaf** 150 ± 30 -28.4 283-Post BP 0 161 N/A N/A

8 Cariac-09-06_6 BA 46 SBOF 1230 ± 30 -20.4 868-691 768 403-54 224

9 Cariac-09-06_7 BA 52 Ptp 610 ± 30 N/A 309-141 258 N/A N/A

10 Cariac-09-06_8 BA 52 SBOF 1210 ± 30 -19.7 834-671 756 356-Post BP 0 199

11 Cariac-09-06_9 BA 65.5 BS 670 ± 30 NA 405-262 316 N/A N/A

12 Cariac-09-06_10 BA 65.5 SBOF 1210 ± 30 -19.7 834-671 737 336-Post BP 0 199

13 Cariac-09-06b CEA-S 80 SBOF 1100 ± 30 -20.3 723-609 659 N/A N/A

14 Cariac-09-06_11 BA 83.5 SBOF 1620 ± 30 -21.0 1263-1101 1191 N/A N/A

15 Cariac-09-06c CEA-S 94 SBOF 1490 ± 30 -20.3 1136-953 1036 N/A N/A

16 Cariac-09-06_12 BA 95 SBOF 1950 ± 30 -22.3 1586-1403 1502 N/A N/A

17 Cariac-09-06_13 BA 101 BS 1000 ± 30 +0.2 640-527 583 N/A N/A

18 Cariac-09-06_14 BA 101 SBOF 1670 ± 30 -19.2 1288-1168 1234 716-505 607

19 Cariac-09-06_15 BA 107 Ptp 1100 ± 30 N/A 723-609 658 N/A N/A

20 Cariac-09-06_16 BA 107 SBOF 1760 ± 30 -20.4 1373-1253 1306 826-547 685

21 Cariac0905B_30 BA 30 SBOF 1040 ± 30 -19.9 661-546 613 226 - Post BP 0 72

22 Cariac09_05B_61 CEA-S 61 SBOF 1390 ± 30 -20.1 1032-875 936 501-276 391

23 Cariac09_05B_91 CEA-S 91 SBOF 1700 ± 30 -19.7 1312-1182 1259 749-518 631

24 Cariac-09-05B CEA-S 95 SBOF 1820 ± 30 -20.1 1460-1285 1356 891-633 739

1
Abbreviations

BA: Beta Analytic; CEA-S: Saclay CEA Laboratory of the CNRS-INSU ARTEMIS; SBOF:

sediment bulk organic fraction); Ptp: Pteropod.


2
We used a range calibration of 95%
*
Age reported in percentage of modern carbon
**
Sample calibrated with the curve INTcal13
Table_2

Table 1. ΔR calculation.
Depth Conventional
Sample Type ΔR (yr)
(cm) age14C BP

Cariac-09-06_3 (388630) 28.5 Bivalve shell 400 ± 30 680 ± 42


Cariac-09-06_4 (388631) SBOF 1080 ± 30
Cariac-09-06_5 (388632) leaf* 536 ± 80
46 694 ± 85
Cariac-09-06_6 (388633) SBOF 1230 ± 30
Cariac-09-06_7 (388634) Ptéropode shell 610 ± 30
52 600 ± 42
Cariac-09-06_8 (388635) SBOF 1210 ± 30
Cariac-09-06_9 (388636) SBOF 670 ± 30
65.5 540 ± 42
Cariac-09-06_10(388637) SBOF 1210 ± 30
Cariac-09-06_13(388640) Bivalve shell 1000 ± 30
101 670 ± 42
Cariac-09-06_14(388641) SBOF 1670 ± 30
Cariac-09-06_15(388642) Ptéropode shell 1100 ± 30
107 660 ± 42
Cariac-09-06_16(388643) SBOF 1760 ± 30
Weighted mean ΔR 633 ± 24
StDev on the ΔR 59
*The reported age corresponds to its marine equivalent.
Figure captions
Click here to download Supplementary Data: Figure captions.docx

You might also like