[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (273)

Search Parameters:
Keywords = Eocene

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 10333 KiB  
Article
How to Recognize Mosses from Extant Groups among Paleozoic and Mesozoic Fossils
by Michael S. Ignatov, Tatyana V. Voronkova, Ulyana N. Spirina and Svetlana V. Polevova
Diversity 2024, 16(10), 622; https://doi.org/10.3390/d16100622 - 8 Oct 2024
Viewed by 607
Abstract
This paper describes a range of Paleozoic and Mesozoic mosses and assesses how far they can be referred to extant taxa at the family, ordinal, or class levels. The present study provides new data on Paleozoic mosses of the order Protosphagnales, re-evaluating affinities [...] Read more.
This paper describes a range of Paleozoic and Mesozoic mosses and assesses how far they can be referred to extant taxa at the family, ordinal, or class levels. The present study provides new data on Paleozoic mosses of the order Protosphagnales, re-evaluating affinities of some groups previously thought to be unrelated. The leaf areolation pattern, combined with the leaf costa anatomy, results in the subdivision of Protosphagnales into five separate families: Protosphagnaceae (at least six genera), Polyssaieviaceae (at least three genera), and three monogeneric families: Rhizonigeritaceae, Palaeosphagnaceae, and Servicktiaceae. We urge caution in referring Paleozoic and Early Mesozoic fossil mosses as members of Dicranidae and Bryidae, as they may belong to the extinct moss order Protosphagnales. Additional evidence supports the relation of the Permian genus Arvildia to extant Andreaeopsida. We segregate Late Palaeozoic and Early Mesozoic mosses that are superficially similar to extant members of either Dicranales or Polytrichales, into the artificial informal group of Archaeodicranids, distinguishing them from ecostate Paleozoic and Mesozoic mosses, which are combined here into another artificial informal group, Bryokhutuliinids. The latter includes the genus Bryokhutuliinia, widespread in contemporary Asia, from the Middle Jurassic to the Lower Cretaceous, as well as other superficially similar ecostate plants from different regions worldwide, ranging from the Upper Palaeozoic to the Lower Cretaceous. A list of Paleozoic, Mesozoic, and Eocene moss fossils suitable for age calibration in phylogenetic trees is provided. Full article
(This article belongs to the Section Plant Diversity)
Show Figures

Figure 1

Figure 1
<p>Protosphagnalean mosses showing typical dimorphic areolation pattern for <span class="html-italic">Protosphagnum</span> (<b>H</b>,<b>I</b>), mostly monomorphic for <span class="html-italic">Intia</span> (<b>B</b>,<b>D</b>,<b>F</b>,<b>J</b>) and <span class="html-italic">Kosjunia</span> (<b>G</b>), and combining mono and dimorphic areolation types in different parts of leaves (<b>A</b>,<b>C</b>,<b>E</b>,<b>H</b>). (<b>A</b>) Stem with leaves, (<b>B</b>) young leaves crowded at stem tip, and (<b>C</b>–<b>J</b>) leaf fragments. Aristovo, Permian (Lopingian): (<b>A</b>) 126A-3A, (<b>B</b>) 49B-5, (<b>C</b>) 16B-6, (<b>D</b>) 39B-7, (<b>E</b>) 100A-9, (<b>F</b>) 44A-1, (<b>G</b>) 47B-9, (<b>H</b>) 31A-3, (<b>I</b>) 38A-6, and (<b>J</b>) 43B-9.</p>
Full article ">Figure 2
<p><span class="html-italic">Protosphagnum nervatum</span>: (<b>A</b>) fully developed apical areolation; (<b>B</b>,<b>D</b>) developing apical areolation; (<b>C</b>,<b>E</b>) young leaves with costa branches confluent with rows of laminal areolation; (<b>F</b>–<b>K</b>) branch primordia with surrounding leaves, red arrow in H points one of primordia magnified in G, LM (<b>F</b>–<b>J</b>) and SEM (<b>K</b>) images. Aristovo, Permian (Lopingian): (<b>A</b>) 5A-3, (<b>B</b>) 11B-2, (<b>C</b>) 19A-2, (<b>D</b>) 47A-1, (<b>E</b>) 19A-3, (<b>F</b>) 126B-1, (<b>G</b>) 126B-2, (<b>H</b>) 126B-2, (<b>I</b>) 126B-4, (<b>J</b>) 126B-1, and (<b>K</b>) CUT_SEM_4.</p>
Full article ">Figure 3
<p><span class="html-italic">Protosphagnum nervatum</span>. (<b>A</b>–<b>C</b>) Araldite-embedded leaf transverse sections, 2 μm thick, under LM (<b>A</b>) and 60 nm thick under TEM (<b>B</b>–<b>C</b>), showing the homogeneous structure of the fossil; (<b>D</b>,<b>E</b>) costa surface; (<b>D</b>–<b>J</b>,<b>L</b>,<b>M</b>) SEM images of leaves, showing dimorphic cells with very thin walls of the hyalocysts (<b>F</b>–<b>J</b>,<b>L</b>,<b>M</b>), partly broken (<b>G</b>,<b>H</b>,<b>J</b>) or mostly retained (<b>I</b>,<b>L</b>,<b>M</b>); (<b>K</b>) juvenile leaf (shown in whole in inset) areolation, LM image, showing the delicate nature of the hyalocyst cell walls. Unistratose part of costa is seen in (<b>I</b>) (arrowed). Aristovo, Permian (Lopingian): (<b>A</b>) CUT_S5_3_37, (<b>B</b>–<b>C</b>) CUT_TEM_5, (<b>D</b>–<b>J</b>,<b>L</b>,<b>M</b>) CUT_SEM_1, and (<b>K</b>) 100B-2.</p>
Full article ">Figure 4
<p><span class="html-italic">Protosphagnum nervatum</span>. (<b>A</b>,<b>B</b>) Transverse costa sections, SEM images, of leaf shown in (<b>D</b>,<b>E</b>), LM images; (<b>C</b>,<b>F</b>–<b>J</b>) transverse costa sections, SEM images, of leaf shown in (<b>H</b>,<b>I</b>), LM images. Note the unistratose costa, in which cells look to be filled with spongy material (<b>C</b>,<b>F</b>), likely an effect of fossilization. Aristovo, Permian (Lopingian): (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) CUT_SEM_2, and (<b>C</b>,<b>F</b>–<b>J</b>) CUT_SEM_3.</p>
Full article ">Figure 5
<p><span class="html-italic">Rhizinigerites neuburgae</span>. (<b>A</b>, with close up of rhizoidophore in <b>G</b>) habit, showing the branched stem with leaves and rhizoidophore; the A-inset shows the branch bud (where the arrow points) that has no foliate structure on the stem around it; the G-inset shows the end of the rhizoidophore with rhizoid clusters. (<b>B</b>) A leaf fragment showing areolation with some of the cells missing and no marginal border; the inset highlights the protosphagnalean areolation pattern. (<b>C</b>,<b>D</b>) The leaf apical parts and areolation in different parts of the leaves. (<b>E</b>,<b>F</b>) The laminal cells, showing the areolation variation. (<b>H</b>) A part of the rhizoidophore separated by places with abundant rhizoids. (<b>I</b>) A rhizoid cluster on the rhizoidophore. (<b>J</b>,<b>K</b>) Lower leaf parts showing the unistratose veins of long cells, diverging from the main costa (red arrows). Viled, Permian (Lopingian). See details of the locality in [<a href="#B15-diversity-16-00622" class="html-bibr">15</a>,<a href="#B16-diversity-16-00622" class="html-bibr">16</a>]. (<b>A</b>–<b>C</b>,<b>E</b>,<b>F</b>,<b>H</b>–<b>K</b>) GIN 3774/3B-10-1, (<b>D</b>) GIN 3774/3B-10-2, and (<b>G</b>) GIN 3774/3B-5-9.</p>
Full article ">Figure 6
<p><span class="html-italic">Palaeosphagnum meyenii</span>. (<b>A</b>–<b>D</b>) Leaf fragments and details of areolation. (<b>E</b>–<b>G</b>) Leaf fragment used for sectioning, shown in (<b>H</b>–<b>K</b>). (<b>H</b>–<b>K</b>) Transverse sections of leaf fragment F, which is 60 nm thick, TEM, showing a unistratose lamina with partly inflated cells (<b>H</b>,<b>J</b>). Multistratose costa (<b>I</b>), and bistratose area flanking the costa (<b>K</b>). Aristovo, Permian (Lopingian), (<b>A</b>) 124B-13, (<b>B</b>) 105B-8, (<b>C</b>) 125A-1, (<b>D</b>) 105B-8, and (<b>E</b>–<b>K</b>) CUT_TEM_P9.</p>
Full article ">Figure 7
<p><span class="html-italic">Servicktia undulata</span>. (<b>A</b>,<b>B</b>) Leaf fragment and the cells of its border. (<b>C</b>–<b>E</b>) Its transverse sections, showing a rough cell surface due to irregular papillae and probably prorate cell ends (<b>C</b>), unistratose lamina (<b>D</b>), and multistratose costa (<b>E</b>). Aristovo, Permian (Wushiapingian), (<b>A</b>–<b>E</b>) CUT_TEM_P11.</p>
Full article ">Figure 8
<p><span class="html-italic">Servicktia vorcutannularioides</span>. (<b>A</b>–<b>D</b>) Leaf fragments and details of dimorphic cell areolation. (<b>E</b>–<b>G</b>) Transverse sections under TEM (<b>E</b>,<b>F</b>) and LM (<b>G</b>), showing multistratose costa (<b>E</b>,<b>G</b>), unistratose lamina (<b>F</b>), and inflated cells on ventral side of costa (E, G). Aristovo, Permian (Lopingian), (<b>A</b>) 107A-2, (<b>B</b>) 106A-5, (<b>C</b>) 106A-5, (<b>D</b>) 106A-10, and (<b>E</b>–<b>G</b>) CUT_TEM_P6.</p>
Full article ">Figure 9
<p><span class="html-italic">Servicktia tatyanae</span>. Holotype: (<b>A</b>–<b>C</b>), whole leaf fragment and close ups of upper and lower leaf parts. Aristovo, Permian (Lopingian), (<b>A</b>–<b>C</b>) 100B-1.</p>
Full article ">Figure 10
<p><span class="html-italic">Polyssaievia spinulifolia</span> (<b>A</b>–<b>H</b>), <span class="html-italic">Polyssaievia deflexa</span> (<b>I</b>), and young leaf of <span class="html-italic">Protosphagnum nervatum</span> (<b>J</b>), showing variation in areolation in different parts of leaves of <span class="html-italic">Polyssaievia</span> and similarity in areolation to juvenile leaf of <span class="html-italic">P. nervatum</span>. (<b>A</b>) shoot, (<b>B</b>–<b>I</b>) leaf fragments, showing ‘net venation’ in proximal part of leaves and prosenchymatous cells in their distal part (<b>B</b>,<b>E</b>,<b>F</b>). (<b>J</b>) Juvenile leaf. Permian (Lopingian) specimens from localities described for A–I in [<a href="#B11-diversity-16-00622" class="html-bibr">11</a>] and for J in [<a href="#B23-diversity-16-00622" class="html-bibr">23</a>]: (<b>A</b>–<b>D</b>) Tunguska coal basin GIN 3087/1019-3, (<b>E</b>,<b>F</b>) Tunguska coal basin GIN 3087/1018-4, (<b>G</b>,<b>H</b>) Kuznetsk coal basin GIN 3026/95A, (<b>I</b>) Pechora coal basin GIN 3041_151c, and (<b>J</b>) Pechora coal basin MHA: Adzva_32M_20_4_A_3.</p>
Full article ">Figure 11
<p><span class="html-italic">Arvildia elenae</span> (<b>A</b>,<b>B</b>,<b>E</b>–<b>K</b>), compared with extant <span class="html-italic">Andreaea rothii</span> F. Weber &amp; D. Mohr (<b>C</b>) and <span class="html-italic">Andreaeobryum macrosporum</span> Steere &amp; B.M. Murray (<b>D</b>,<b>L</b>). (<b>A</b>,<b>B</b>) Leaf apical part showing uniseriate and then, from the 5th cell from the apex, abruptly a biseriate cell arrangement, typical of the Andreaeales and Andreaeobryales. (<b>C</b>,<b>D</b>) Leaf apices with uniseriate and then biseriate cells. (<b>E</b>,<b>F</b>) Apical parts of two leaves, apparently from the same shoot, and their transverse sections showing the most distal unicellular part of the leaf on the left, as shown in the 2 μm section under LM. (<b>G</b>,<b>H</b>) A densely foliate shoot and sublongitudinal section of the apical leaf part under TEM. (<b>I</b>–<b>K</b>) A leaf and its cross-section under TEM, showing only moderately differentiated costal cells, comparable to the <span class="html-italic">Andreaeobryum</span> sections, shown in L. (<b>L</b>) Transverse sections of the distal leaf parts above the stem apex of <span class="html-italic">Andreaeobryum</span>, showing a moderately differentiated costa, a 2 μm section, with fluorescent microscopy. <span class="html-italic">Arvildia</span> collections are from Aristovo, Permian (Lopingian): (<b>A</b>,<b>B</b>) 113-1, (<b>E</b>,<b>F</b>) CUT_TEM_P15, (<b>G</b>,<b>H</b>) CUT_TEM_P16, and (I–K) CUT_TEM_P12. Extant specimens are from: (<b>C</b>) Norway, MW9000070, and (<b>D</b>,<b>L</b>) Russia, MHA9022375.</p>
Full article ">Figure 12
<p><span class="html-italic">Kulindobryum taylorioides</span> (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) compared with extant <span class="html-italic">Tayloria splachnoides</span> (Schleich. ex Schwär.) Hook. (<b>C</b>). ((<b>A</b>), and its close up in (<b>B</b>)): capsule with long neck, partly broken at mouth, where pendent peristome teeth occur, (<b>C</b>) open capsule with 32 peristome teeth, (<b>D</b>) obliquely compressed capsule showing deoperculate mouth with peristome teeth fragments, and (<b>E</b>) still operculate capsule covered by calyptra. Peristome teeth or their fragments arrowed. Transbaikalia, Kulinda, Middle Jurassic (for locality information see references in [<a href="#B18-diversity-16-00622" class="html-bibr">18</a>]): ((<b>A</b>),(<b>B</b>): PIN 5648/2, (<b>D</b>): PIN 5648/3, (<b>E</b>): PIN 5648/1) and (<b>C</b>) Russia, Urals, MHA9020838.</p>
Full article ">
34 pages, 22660 KiB  
Article
Source Rock Evaluation and Hydrocarbon Expulsion Characteristics of Effective Source Rocks in the Fushan Depression, Beibuwan Basin, China
by Xirong Wang, Fujie Jiang, Xiaowei Zheng, Di Chen, Zhenguo Qi, Yilin Liu, Jing Guo and Yuqi Zhang
Minerals 2024, 14(10), 975; https://doi.org/10.3390/min14100975 - 27 Sep 2024
Viewed by 328
Abstract
This study presents an integrated approach using organic geochemistry and incident-light organic petrographic microscopy techniques to characterize the kerogen type, hydrocarbon potential, thermal maturity, and effective depositional environment of the Eocene Liushagang Formation intervals in the western Huangtong Sag, eastern Bailian Sag, central [...] Read more.
This study presents an integrated approach using organic geochemistry and incident-light organic petrographic microscopy techniques to characterize the kerogen type, hydrocarbon potential, thermal maturity, and effective depositional environment of the Eocene Liushagang Formation intervals in the western Huangtong Sag, eastern Bailian Sag, central Huachang Sub-uplift, and Southern Slope Zone area in the Fushan Depression, Beibuwan Basin. The results show that the hydrocarbon potential of these organic-rich lacustrine shale areas is mainly dependent on the depositional environment and the present-day burial depth of sediments. Oscillations and transitions between (i) rocks with dominant allochthonous organic matter (including primary/reworked vitrinite and inertinite macerals and terrestrial debris particles) representing a large influence of continental sediments (e.g., source supply direction) and (ii) rocks with dominant autochthonous organic matter (e.g., alginite) indicate a distal and stable lacustrine basin depositional environment. The source rock thickness ranges from 40.1 to 387.4 m. The average TOC of the Liushagang Formation in the Fushan Sag is between 0.98% and 2.00%, with the highest organic matter abundance being in the first and second sections of the Liushagang Formation, presenting as high-quality source rocks. The organic matter is predominantly Type II1 and Type II2. The highest vitrinite reflectance (1.14%) is in the Huangtong and Bailian Sags. The source rocks of the second section of the Liushagang Formation are primary hydrocarbon generators, contributing 55.11% of the total generation. Hydrocarbon sequestration peaks at %Ro 0.80%, with a maximum efficiency of 97.7%. The cumulative hydrocarbon generation of the Liushagang Formation is 134.10 × 108 tons, with 50.52 × 108 tons having been expelled and 83.58 × 108 tons remaining. E2L2X and E2L2S have maximum hydrocarbon displacement intensities of 184.22 × 104 t/km² and 45.39 × 104 t/km², respectively, with cumulative displacements of 52.99 × 108 tons and 15.58 × 108 tons. The oil and gas accumulation system is highly prospective, showing significant exploration potential. Full article
(This article belongs to the Section Mineral Exploration Methods and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Beibuwan Basin. (<b>b</b>) Location of the Fushan Depression (modified after [<a href="#B56-minerals-14-00975" class="html-bibr">56</a>]). (<b>c</b>) Location of western Huangtong Sag, eastern Bailian Sag, central Huachang Sub-uplift, and Southern Slope Belt in the Fushan Depression (modified after [<a href="#B57-minerals-14-00975" class="html-bibr">57</a>]).</p>
Full article ">Figure 2
<p>Stratigraphic column of the Fushan Depression (modified after [<a href="#B13-minerals-14-00975" class="html-bibr">13</a>]). E<sub>2</sub>L<sub>1</sub>, E<sub>2</sub>L<sub>2</sub>, and E<sub>2</sub>L<sub>3</sub> are Members 1, 2, and 3 of the Eocne Liushagang Formation. According to the lithological characteristics and the production demand of China‘s oil fields, it is divided into the E<sub>2</sub>L<sub>3X</sub>, E<sub>2</sub>L<sub>3Z</sub>, E<sub>2</sub>L<sub>3S</sub>, E<sub>2</sub>L<sub>2X</sub>, E<sub>2</sub>L<sub>2S</sub>, E<sub>2</sub>L<sub>1X</sub>, and E<sub>2</sub>L<sub>1S</sub> sequentially from the top to bottom of Eocene source rocks.</p>
Full article ">Figure 3
<p>Sample position distribution in the Liushagang Formation in Fushan Depression. The wells in the west of the study area, such as <span class="html-italic">FC1</span>, were assigned to the Huantong Sag. The <span class="html-italic">L</span> well in the east belongs to the Bailian area, which is assigned to the Bailian Sag. The wells in the central turning zone north of the Meihua Fault are classified into the Huachang Sub-uplift, and the wells north of the Meihua Fault are located in the Southern Slope Belt.</p>
Full article ">Figure 4
<p>Relationship between total organic carbon content and hydrocarbon generation potential of source rocks in the Huangtong Sag, Huachang Sub-uplift, Bailian Sag, and Southern Slope Belt in the Fushan Depression. The E<sub>2</sub>L<sub>2X</sub> stratigraphic sub-section exhibits superior geological parameters compared with its counterparts, indicating significant hydrocarbon generation potential. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks (after [<a href="#B74-minerals-14-00975" class="html-bibr">74</a>]).</p>
Full article ">Figure 5
<p>TOC contour map of the Liuliushagang Formation in the Fushan Depression. The TOC content ranges from 0.5% to 3.6%, indicating a substantial abundance of organic matter. The elevated concentrations are predominantly found within the Huangtong and Bailian Sag regions, with a gradual decrease towards the peripheral areas. The E<sub>2</sub>L<sub>2X</sub> and E<sub>2</sub>L<sub>2S</sub> strata exhibit superior characteristics compared with the other geological layers. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 5 Cont.
<p>TOC contour map of the Liuliushagang Formation in the Fushan Depression. The TOC content ranges from 0.5% to 3.6%, indicating a substantial abundance of organic matter. The elevated concentrations are predominantly found within the Huangtong and Bailian Sag regions, with a gradual decrease towards the peripheral areas. The E<sub>2</sub>L<sub>2X</sub> and E<sub>2</sub>L<sub>2S</sub> strata exhibit superior characteristics compared with the other geological layers. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 6
<p>Extracted bitumen <span class="html-italic">A</span> contour map of the Liushagang Formation in the Fushan Depression. The extracted bitumen A content ranges from 0.01% to 0.54%, indicating a substantial organic matter abundance. The highest concentrations are predominantly found within the Bailian Sag, while the Huachang Sub-uplift exhibits elevated values in the E<sub>2</sub>L<sub>2X</sub> and E<sub>2</sub>L<sub>2S</sub> intervals, as well as in the E<sub>2</sub>L<sub>1X</sub> stratigraphic unit. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 6 Cont.
<p>Extracted bitumen <span class="html-italic">A</span> contour map of the Liushagang Formation in the Fushan Depression. The extracted bitumen A content ranges from 0.01% to 0.54%, indicating a substantial organic matter abundance. The highest concentrations are predominantly found within the Bailian Sag, while the Huachang Sub-uplift exhibits elevated values in the E<sub>2</sub>L<sub>2X</sub> and E<sub>2</sub>L<sub>2S</sub> intervals, as well as in the E<sub>2</sub>L<sub>1X</sub> stratigraphic unit. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 7
<p>The relationship between HI and Tmax of source rocks in the Huangtong sag, Huachang Sub-uplift, Bailian Sag, and Southern Slope Belt in the Fushan Depression. The predominant type of organic matter is classified as type II<sub>2</sub>. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks. Thermal maturity of organic matter (after [<a href="#B24-minerals-14-00975" class="html-bibr">24</a>]).</p>
Full article ">Figure 8
<p>Variation in the vitrinite reflectance of source rocks of E<sub>2</sub>L<sub>3X</sub>, E<sub>2</sub>L<sub>3Z</sub>, E<sub>2</sub>L<sub>3S</sub>, E<sub>2</sub>L<sub>2X</sub>, E<sub>2</sub>L<sub>2S</sub>, E<sub>2</sub>L<sub>1X</sub>, and E<sub>2</sub>L<sub>1S</sub> sequentially with depth in the Fushan Depression. (<b>a</b>) Huangtong Sag, (<b>b</b>) Huachang Uplift, (<b>c</b>) Bailian Sag, and (<b>d</b>) Southern Slope Belt. The overall variation in the Rock-Eval pyrolysis parameter, <span class="html-italic">Ro</span>, exhibits a positive correlation with the increase in burial depth.</p>
Full article ">Figure 9
<p>Ro plane contour map of source rocks in the Liushagang Formation in the Fushan Depression. The region of elevated Ro values is situated within the Huangtong Sag and the Bailian Sag. There is a gradual decrease in Ro values from these high-value areas towards the periphery of the respective depressions. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 9 Cont.
<p>Ro plane contour map of source rocks in the Liushagang Formation in the Fushan Depression. The region of elevated Ro values is situated within the Huangtong Sag and the Bailian Sag. There is a gradual decrease in Ro values from these high-value areas towards the periphery of the respective depressions. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 10
<p>Source rock thickness contour map of the Liushagang Formation in the Fushan Depression. Influenced by the Meihua Fault, the provenance rocks of the Early Liushagang Formation in the northern region underwent significant denudation. The depositional center and the area of maximum thickness of these source rocks are observed in the Huangtong Sag and the Bailian Sag, respectively. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 10 Cont.
<p>Source rock thickness contour map of the Liushagang Formation in the Fushan Depression. Influenced by the Meihua Fault, the provenance rocks of the Early Liushagang Formation in the northern region underwent significant denudation. The depositional center and the area of maximum thickness of these source rocks are observed in the Huangtong Sag and the Bailian Sag, respectively. (<b>a</b>) E<sub>2</sub>L<sub>3X</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>3Z</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>3S</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>2X</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2S</sub> source rocks; (<b>f</b>) E<sub>2</sub>L<sub>1X</sub> source rocks; and (<b>g</b>) E<sub>2</sub>L<sub>1S</sub> source rocks.</p>
Full article ">Figure 11
<p>Triangle diagram of macerals of organic matter in the source rocks of the Liushagang Formation in the Fushan Depression. The sapropelic and exinitic maceral groups exhibit elevated concentrations, with values ranging from 62.7% to 87.4% and an average of 72.57%. As the principal constituents responsible for hydrocarbon generation, this high content underscores the robust hydrocarbon-generating potential of the organic matter.</p>
Full article ">Figure 12
<p>Microscopic identification of kerogen macerals and types of the Liushagang Formation in the Fushan Depression. (<b>a</b>) Y13X in the Huangtong Sag, E<sub>2</sub>L<sub>1</sub>, 3586.7 m, type III kerogen. The inertinite macerals can be observed, and (<b>b</b>) the sapropel group is not strongly reflected under fluorescence. (<b>c</b>) H7X in the Huachang Sub-uplift, E<sub>2</sub>L<sub>1</sub>, 2401.62 m, type III kerogen. (<b>d</b>) The vitrinite macerals can be observed. (<b>e</b>) Y7X in the Huangtong Sag, E<sub>2</sub>L<sub>1</sub>, 3744.12 m, type II<sub>2</sub> kerogen; (<b>f</b>) Development of sapropelic macerals. (<b>g</b>) H7X in the Huachang Sub-uplift, 3448.67 m, E<sub>2</sub>L<sub>3</sub>, type II<sub>2</sub> kerogen. (<b>h</b>) Compared with E<sub>2</sub>L<sub>1</sub>, more sapropelic macerals were developed.</p>
Full article ">Figure 13
<p>Sedimentary burial history and hydrocarbon generation history. (<b>a</b>) The Y7 well in the Huangtong Sag; (<b>b</b>) the L23 well in the Bailian Sag. The Huangtong Sag and the Bailian Sag predominantly accumulated hydrocarbons during the E<sub>2</sub>L<sub>2</sub> and E<sub>2</sub>L<sub>1</sub> epochs following the generation of hydrocarbons, indicative of an early phase of two-stage accumulation. Subsequently, the hydrocarbons migrated to the structural high areas, namely, the Huachang Sub-uplift and the Southern Slope Belt, marking a later phase characterized by one-stage accumulation.</p>
Full article ">Figure 14
<p>Hydrocarbon expulsion model of the Liushagang Formation in the Fushan Depression, the red spot is the hydrocarbon expulsion threshold to restore the maximum hydrocarbon generation potential index. Upon reaching the critical depths where the source rock crosses the thresholds for hydrocarbon generation and expulsion, the rate of hydrocarbon generation peaks at a depth of 2370 m. Conversely, at a depth of 3220 m, the expulsion rate of hydrocarbons is maximized, achieving a peak expulsion efficiency of 88%.</p>
Full article ">Figure 15
<p>Contour map of hydrocarbon generation and expulsion intensity of source rocks in each layer of the Liushagang Formation in the Fushan Depression. (<b>a</b>) E<sub>2</sub>L<sub>3</sub> source rocks; (<b>b</b>) E<sub>2</sub>L<sub>2</sub> source rocks; (<b>c</b>) E<sub>2</sub>L<sub>1</sub> source rocks; (<b>d</b>) E<sub>2</sub>L<sub>3</sub> source rocks; (<b>e</b>) E<sub>2</sub>L<sub>2</sub> source rocks; and (<b>f</b>) E<sub>2</sub>L<sub>1</sub> source rocks.</p>
Full article ">Figure 16
<p>(<b>a</b>) Isoline diagram of hydrocarbon generation of the source rocks in the Liushagang Formation in the Fushan Depression. (<b>b</b>) Isoline diagram of expulsion intensity of the source rocks in the Liushagang Formation in the Fushan Depression. The E<sub>2</sub>L<sub>2</sub> dark mudstone exhibits a higher hydrocarbon generation potential compared with the other strata. It shares similar characteristics in hydrocarbon expulsion and generation processes. In contrast, the contributions from the E<sub>2</sub>L<sub>3</sub> and E<sub>2</sub>L<sub>1</sub> source rocks are comparatively minor.</p>
Full article ">
47 pages, 19713 KiB  
Article
Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon
by Nasser Farhat
Hydrology 2024, 11(9), 156; https://doi.org/10.3390/hydrology11090156 - 21 Sep 2024
Viewed by 1084
Abstract
Countries face challenges of excess, scarcity, pollution, and uneven water distribution. This study highlights the benefits of advances in groundwater engineering that improve the understanding of utilizing local geological characteristics due to their crucial role in resisting drought in southern Lebanon. The type [...] Read more.
Countries face challenges of excess, scarcity, pollution, and uneven water distribution. This study highlights the benefits of advances in groundwater engineering that improve the understanding of utilizing local geological characteristics due to their crucial role in resisting drought in southern Lebanon. The type of drought in the region was determined using the Standardized Precipitation Index (SPI), Standardized Vegetation Index (NDVI), Vegetation Condition Index (VCI), and Soil Moisture Anomaly Index (SM). The dry aquifer and its characteristics were analyzed using mathematical equations and established hydrogeological principles, including Darcy’s law. Additionally, a morphometric assessment of the Litani River was performed to evaluate its suitability for artificial recharge, where the optimal placement of the water barrier and recharge tunnels was determined using Spearman’s rank correlation coefficient. This analysis involved excluding certain parameters based on the Shapiro–Wilk test for normality. Accordingly, using the Geographic Information System (GIS), we modeled and simulated the potential water table. The results showed the importance and validity of linking groundwater engineering and morphometric characteristics in combating the drought of groundwater layers. The Eocene layer showed a clearer trend for the possibility of being artificially recharged from the Litani River than any other layer. The results showed that the proposed method can enhance artificial recharge, raise the groundwater level to four levels, and transform it into a large, saturated thickness. On the other hand, it was noted that the groundwater levels near the surface will cover most of the area of the studied region and could potentially store more than one billion cubic meters of water, mitigating the effects of climate change for decades. Full article
(This article belongs to the Section Surface Waters and Groundwaters)
Show Figures

Figure 1

Figure 1
<p>The study area in Lebanon.</p>
Full article ">Figure 2
<p>Annual rainfall in the study area.</p>
Full article ">Figure 3
<p>Trend in temperature increase.</p>
Full article ">Figure 4
<p>The lithological map of the study area.</p>
Full article ">Figure 5
<p>A geological section from east to west showing the fractures and the inclination of the geological strata in addition to the Bint Jbeil syncline. E2: Eocene; C6: Senonian; C4: Cenomanian.</p>
Full article ">Figure 6
<p>Stratum with the lineaments and faults in the study area.</p>
Full article ">Figure 7
<p>The topography of the region.</p>
Full article ">Figure 8
<p>The Standardized Precipitation Index (SPI) from 1981 to 2024.</p>
Full article ">Figure 9
<p>Vegetation Condition Index (VCI).</p>
Full article ">Figure 10
<p>Soil Moisture Anomaly Index (SM).</p>
Full article ">Figure 11
<p>DEM with dimension of the Eocene aquifer.</p>
Full article ">Figure 12
<p>Geological section of the syncline of the Eocene aquifer. C6 is the Senonian isolate layer; C4 is the Cenomanian aquifer; E2 is the Eocene saturated thickness.</p>
Full article ">Figure 13
<p>Aquifers in the study area.</p>
Full article ">Figure 14
<p>Hydrological map of the study area.</p>
Full article ">Figure 15
<p>Litani watershed in the study area.</p>
Full article ">Figure 16
<p>The Litani River in its valley and the optimal site of the dam.</p>
Full article ">Figure 16 Cont.
<p>The Litani River in its valley and the optimal site of the dam.</p>
Full article ">Figure 17
<p>The water table in the Eocene aquifer showing its relationship with the proposed tunnels.</p>
Full article ">Figure 18
<p>Topographic section from the dam to the southern part of the study area showing the tunnels and the water table.</p>
Full article ">Figure 19
<p>Regional depth of the water table following filling of the Litani River reservoir.</p>
Full article ">Figure 20
<p>This figure shows the potential groundwater level in a very small area within deep valleys where the water table is theoretically supposed to approach the ground surface.</p>
Full article ">Figure 21
<p>Areas where the water table will be at depths ranging from 1 to 150 m.</p>
Full article ">Figure 22
<p>Areas where the water table will be at depths ranging from 150 to 250 m.</p>
Full article ">Figure 23
<p>Areas where the water table will be at depths greater than 250 m.</p>
Full article ">Figure 24
<p>Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (<b>A</b>) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (<b>B</b>) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (<b>C</b>) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (<b>D</b>) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (<b>E</b>) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (<b>F</b>) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (<b>G</b>) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.</p>
Full article ">Figure 24 Cont.
<p>Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (<b>A</b>) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (<b>B</b>) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (<b>C</b>) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (<b>D</b>) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (<b>E</b>) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (<b>F</b>) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (<b>G</b>) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.</p>
Full article ">
21 pages, 10207 KiB  
Article
Hydrothermal Karstification of the Pre-Messinian Eonile Canyon: Geomorphological and Geochemical Evidences for Hypogene Speleogenesis in the Middle Nile Valley of Egypt
by Ashraf A. Mostafa, Hatem M. El-Desoky, Diaa A. Saadawi, Ahmed M. Abdel-Rahman, John Webb, Hassan Alzahrani, Fahad Alshehri, Abdurraouf Okok, Ahmed E. Khalil and Eman A. Marghani
Minerals 2024, 14(9), 946; https://doi.org/10.3390/min14090946 - 16 Sep 2024
Viewed by 786
Abstract
The surface and subsurface karst features of the Eocene limestone plateaus along the Middle Nile Valley in Egypt were formerly believed to be epigene in origin and to have developed during post-Eocene pluvial periods. However, the morphology of the caves and their restriction [...] Read more.
The surface and subsurface karst features of the Eocene limestone plateaus along the Middle Nile Valley in Egypt were formerly believed to be epigene in origin and to have developed during post-Eocene pluvial periods. However, the morphology of the caves and their restriction to particular stratigraphic intervals suggests that they are hypogene. The geochemistry and mineralogy of the soft, thick-bedded, brown/black cave infills shows that these sediments originated from hydrothermal processes, as evidenced by their Fe, Mn, Co, Ni, and Cu concentrations. Thus, the karst features are hypogene and probably formed during the opening of the Red Sea Rift at the end of the Oligocene and early Miocene. At this time, there was abundant volcanic activity, as shown by basalt lavas ~70 km northwest of Assiut; this triggered the release of large amounts of CO2 that made the hydrothermal waters acidic and dissolved the caves. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the El-Balayza hypogenic caves west of Assiut (shown by yellow dots).</p>
Full article ">Figure 2
<p>(<b>A</b>) The geomorphological main units in the study area. The yellow dashed line indicates the upper limit of the Nile Valley scarp (unit 2); the red arrow points north. (<b>B</b>) Topographical section showing the relation between hypogenic caves and the main geomorphological units in Middle Nile Valley in Assiut.</p>
Full article ">Figure 3
<p>(<b>A</b>) Geologic map of the studied district (modified after [<a href="#B31-minerals-14-00946" class="html-bibr">31</a>]). (<b>B</b>) Simplified stratigraphic sequence showing the different formations in the early Eocene on the western edge of the Nile Valley in the Drunka region (modified after [<a href="#B29-minerals-14-00946" class="html-bibr">29</a>,<a href="#B30-minerals-14-00946" class="html-bibr">30</a>]).</p>
Full article ">Figure 4
<p>(<b>A</b>) The distinctive stratigraphic control on distribution of the hypogene caves on the western side of the Nile Valley. The two yellow lines define the transverse extent of the hypogene caves between the Drunka and Zawiya Formations, (H) indicates the relict hypogene surface (mesa) exposed by erosional retreat of the Drunka Formation above it. (<b>B</b>) Exposed relict hypogenic features. Notice the person standing completely inside one of the hypogenic fractures that fed the hypogenic system. (<b>C</b>) Hypogene channels containing hydrothermal black sediments and minerals. (<b>D</b>) Remnants of the hypogenic channels (yellow arrows) that represent the flow paths of ascending solutions in the Zawiya Formation, exposed within the transitional desert zone. (<b>E</b>) Irregular limestone fragments within breccia zone.</p>
Full article ">Figure 5
<p>Different features of the hypogene caves west of Assiut. (<b>A</b>–<b>C</b>) Cave entrances between the Zawiya and Drunka Formations, showing typical transverse extension. (<b>A</b>,<b>D</b>) Feeder channels, (<b>A</b>) fissure- and rift-like feeders in passage floors, (<b>D</b>) point feeders. (<b>E</b>,<b>F</b>) Small, isolated holes and condensation channels in El-Balayza caves. (<b>F</b>) Outlet cupolas in the passage ceiling.</p>
Full article ">Figure 6
<p>Diagrammatic representation of morphological geothermal features diagnostic of hypogene cave; modified from [<a href="#B9-minerals-14-00946" class="html-bibr">9</a>].</p>
Full article ">Figure 7
<p>Limestone photomicrographs. (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) Echinoid fragments, some with calcite overgrowths. (<b>C</b>) Micrite matrix with recrystallized patches and large void filled with mosaic macrocrystalline calcite. (<b>F</b>) Recrystallized calcite molluscan (gastropod) shell fragments embedded in micrite matrix.</p>
Full article ">Figure 8
<p>Compositional trends in host limestone. (<b>a</b>) CaO wt.% versus Al<sub>2</sub>O<sub>3</sub> wt.%, (<b>b</b>) CaO wt.% versus SiO<sub>2</sub> wt.%, and (<b>c</b>) CaO wt.% versus Fe<sub>2</sub>O<sub>3</sub> wt.%. (<b>d</b>) Al<sub>2</sub>O<sub>3</sub>-CaO-(MgO + Fe<sub>2</sub>O<sub>3</sub>) ternary diagram. (<b>e</b>) Relationship between Mn and Sr content. (<b>f</b>) Ternary diagram of Fe-Mn-(Co + Ni + Cu) × 10 for host limestone (Ca-rich samples) and iron-rich brown/black cave infill (low-Ca samples) [<a href="#B54-minerals-14-00946" class="html-bibr">54</a>].</p>
Full article ">Figure 9
<p>(<b>A</b>) Sediment fill within hypogenic cave, showing layers of yellow-ochre, brown, and black color. (<b>B</b>) Large calcite crystals within void in hydrothermally altered limestone.</p>
Full article ">Figure 10
<p>XRD patterns of samples of cave fill. (<b>A</b>) Pure limestone rich by calcite, quartz and anhydrite. (<b>B</b>) black/brown hand specimen rich by kaolinite and rhodochrosite associated to white calcite. (<b>C</b>) grey color sample rich by magnetite, quartz and piemontite. (<b>D</b>) grey sample rich by iron minerals (geothite) and quartz. (<b>E</b>) milky sample refer to carbonate (dolomite + calcite) and quartz bearing geothite. (<b>F</b>) soft sample bearing witherite.</p>
Full article ">Figure 11
<p>Representative SEM-BSE images showing high peaks of Ca, Fe, Si, Al, Mg, Mn, and Ba, which are represented by calcite, goethite, kaolinite, dolomite, piemontite, and witherite grains, respectively. (<b>A</b>) subhedral crystals rich by quartz carbonate associated to Mn-Fe deposits. (<b>B</b>) fine anhedral crystal bearing CaCO<sub>3</sub> associated to Fe, Si, Al, K, Mn, Fe. (<b>C</b>) anhedral crystal with Ca, Fe, Si, Al, Mg, Mn, and Ba.</p>
Full article ">Figure 12
<p>Distribution of major oxides (<b>A</b>) and trace elements (<b>B</b>) for the iron-rich black/dark brown cave fill samples (upper graphs) and host limestone (lower graphs).</p>
Full article ">
16 pages, 18129 KiB  
Article
Hydraulic and Hydrogeochemical Characterization of Carbonate Aquifers in Arid Regions: A Case from the Western Desert, Egypt
by Mahmoud M. Khalil, Mostafa Mahmoud, Dimitrios E. Alexakis, Dimitra E. Gamvroula, Emad Youssef, Esam El-Sayed, Mohamed H. Farag, Mohamed Ahmed, Peiyue Li, Ahmed Ali and Esam Ismail
Water 2024, 16(18), 2610; https://doi.org/10.3390/w16182610 - 14 Sep 2024
Viewed by 518
Abstract
Using geochemical and pumping test data from 80 groundwater wells, the chemical, hydrologic, and hydraulic properties of the fractured Eocene carbonate aquifer located west of the Al-Minya district, the Western Desert, Egypt, have been characterized and determined to guarantee sustainable management of groundwater [...] Read more.
Using geochemical and pumping test data from 80 groundwater wells, the chemical, hydrologic, and hydraulic properties of the fractured Eocene carbonate aquifer located west of the Al-Minya district, the Western Desert, Egypt, have been characterized and determined to guarantee sustainable management of groundwater resources under large-scale desert reclamation projects. The hydrochemical data show that groundwater from the fractured Eocene carbonate aquifer has a high concentration of Na+ and Cl and varies in salinity from 2176 to 2912 mg/L (brackish water). Water–rock interaction and ion exchange processes are the most dominant processes controlling groundwater composition. The carbonate aquifer exists under confined to semi-confined conditions, and the depth to groundwater increases eastward. From the potentiometric head data, deep-seated faults are the suggested pathways for gas-rich water ascending from the deep Nubian aquifer system into the overlying shallow carbonate aquifer. This mechanism enhances the dissolution and karstification of carbonate rocks, especially in the vicinity of faulted sites, and is supported by the significant loss of mud circulation during well drilling operations. The average estimated hydraulic parameters, based on the analysis of step-drawdown, long-duration pumping and recovery tests, indicate that the Eocene carbonate aquifer has a wide range of transmissivity (T) that is between 336.39 and 389,309.28 m2/d (average: 18,405.21 m2/d), hydraulic conductivity (K) between 1.31 and 1420.84 m/d (average: 70.29 m/d), and specific capacity (Sc) between 44.4 and 17,376.24 m2/d (average: 45.24 m2/d). On the other hand, the performance characteristics of drilled wells show that well efficiency ranges between 0.47 and 97.08%, and well losses range between 2.92 and 99.53%. In addition to variations in carbonate aquifer thickness and clay/shale content, the existence of strong karstification features, i.e., fissures, fractures or caverns, and solution cavities, in the Eocene carbonate aquifer are responsible for variability in the K and T values. The observed high well losses might be related to turbulent flow within and adjacent to the wells drilled in conductive fracture zones. The current approach can be further used to enhance local aquifer models and improve strategies for identifying the most productive zones in similar aquifer systems. Full article
Show Figures

Figure 1

Figure 1
<p>A location map of Egypt showing the distribution of carbonate rocks/potential karst aquifers.</p>
Full article ">Figure 2
<p>(<b>a</b>) A geologic map of the west Al-Minya reclamation project, the Western Desert, Egypt. (<b>b</b>) A location map of 80 wells drilled in the fractured Eocene carbonate aquifer. The two insets show (<b>c</b>) the structural trends of surface faults, in the blue rose diagram in the upper right corner, and (<b>d</b>) deep-seated faults, in the red rose diagram in the lower right corner.</p>
Full article ">Figure 3
<p>Hydrogeologic cross sections along the study area. See <a href="#water-16-02610-f002" class="html-fig">Figure 2</a>b for the cross-section location.</p>
Full article ">Figure 4
<p>A potentiometric head map of the fractured Eocene carbonate aquifer in the study area.</p>
Full article ">Figure 5
<p>Examples of (<b>a</b>) step-drawdown pumping test for well no. 4, (<b>b</b>) long duration or constant-rate pumping test for well no. 22, (<b>c</b>) recovery pumping test for well no.14.</p>
Full article ">Figure 6
<p>(<b>a</b>) A box plot showing the variation of the major ions in the groundwater of the fractured Eocene carbonate aquifer, and (<b>b</b>) a Piper diagram of 30 groundwater samples from the fractured Eocene carbonate aquifer.</p>
Full article ">Figure 7
<p>Binary plots of (<b>a</b>) (Ca<sup>2+</sup> + Mg<sup>2+</sup>) vs. (HCO<sub>3</sub><sup>−</sup> + SO<sub>4</sub><sup>2−</sup>), (<b>b</b>) Ca<sup>2+</sup> vs. HCO<sub>3</sub><sup>−</sup>, (<b>c</b>) Ca<sup>2+</sup> vs. SO<sub>4</sub><sup>2−</sup>, (<b>d</b>) Na<sup>+</sup> + K<sup>+</sup> vs. total cations (TZ<sup>+</sup>), (<b>e</b>) Na<sup>+</sup> vs. Cl<sup>−</sup>, and (<b>f</b>) (Ca<sup>2+</sup> + Mg<sup>2+</sup>) − (SO<sub>4</sub><sup>2−</sup> + HCO<sub>3</sub><sup>−</sup>) vs. (Na<sup>+</sup> + K<sup>+</sup>) − Cl<sup>−</sup>.</p>
Full article ">Figure 8
<p>A spatial distribution map of the (<b>a</b>) specific capacity (Sc), (<b>b</b>) transmissivity (T), and (<b>c</b>) hydraulic conductivity (K) of groundwater wells in the fractured Eocene carbonate aquifer.</p>
Full article ">
21 pages, 9535 KiB  
Article
Petrogenesis of Eocene A-Type Granite Associated with the Yingpanshan–Damanbie Regolith-Hosted Ion-Adsorption Rare Earth Element Deposit in the Tengchong Block, Southwest China
by Zhong Tang, Zewei Pan, Tianxue Ming, Rong Li, Xiaohu He, Hanjie Wen and Wenxiu Yu
Minerals 2024, 14(9), 933; https://doi.org/10.3390/min14090933 - 12 Sep 2024
Viewed by 412
Abstract
The ion-adsorption-type rare earth element (iREE) deposits dominantly supply global resources of the heavy rare earth elements (HREEs), which have a critical role in a variety of advanced technological applications. The initial enrichment of REEs in the parent granites controls the formation of [...] Read more.
The ion-adsorption-type rare earth element (iREE) deposits dominantly supply global resources of the heavy rare earth elements (HREEs), which have a critical role in a variety of advanced technological applications. The initial enrichment of REEs in the parent granites controls the formation of iREE deposits. Many Mesozoic and Cenozoic granites are associated with iREE mineralization in the Tengchong block, Southwest China. However, it is unclear how vital the mineralogical and geochemical characteristics of these granites are to the formation of iREE mineralization. We conducted geochronology, geochemistry, and Hf isotope analyses of the Yingpanshan–Damanbie granitoids associated with the iREE deposit in the Tengchong block with the aims to discuss their petrogenesis and illustrate the process of the initial REE enrichment in the granites. The results showed that the Yingpanshan–Damanbie pluton consists of syenogranite and monzogranite, containing REE-bearing accessory minerals such as monazite, xenotime, apatite, zircon, allanite, and titanite, with a high REE concentration (210–626 ppm, mean value is 402 ppm). The parent granites have Zr + Nb + Ce + Y (333–747 ppm) contents and a high FeOT/MgO ratio (5.89–11.4), and are enriched in Th (mean value of 43.6 ppm), U (mean value of 4.57 ppm), Zr (mean value of 305 ppm), Hf (mean value of 7.94 ppm), Rb (mean value of 198 ppm), K (mean value of 48,902 ppm), and have depletions of Sr (mean value of 188 ppm), Ba (mean value of 699 ppm), P (mean value of 586 ppm), Ti (mean value of 2757 ppm). The granites plot in the A-type area in FeOT/MgO vs. Zr + Nb + Ce + Y and Zr vs. 10,000 Ga/Al diagrams, suggesting that they are A2-type granites. These granites are believed to have formed through the partial melting of amphibolites at a post-collisional extension setting when the Tethys Ocean closed. REE-bearing minerals (e.g., apatite, titanite, allanite, and fluorite) and rock-forming minerals (e.g., potassium feldspar, plagioclase, biotite, muscovite) supply rare earth elements in weathering regolith for the Yingpanshan–Damanbie iREE deposit. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Tectonic map of the eastern Tethys domain (modified after Wang et al. [<a href="#B47-minerals-14-00933" class="html-bibr">47</a>]); (<b>b</b>) Geological map of the Tengchong block with iREE deposits (modified after Deng et al. [<a href="#B48-minerals-14-00933" class="html-bibr">48</a>]); (<b>c</b>) U–Pb ages histogram of zircons from magmatic rocks in the Tengchong block (date from He et al. [<a href="#B15-minerals-14-00933" class="html-bibr">15</a>], Dong et al. [<a href="#B16-minerals-14-00933" class="html-bibr">16</a>], Xu et al. [<a href="#B18-minerals-14-00933" class="html-bibr">18</a>], Yang et al. [<a href="#B19-minerals-14-00933" class="html-bibr">19</a>], Li et al. [<a href="#B26-minerals-14-00933" class="html-bibr">26</a>], Zou et al. [<a href="#B30-minerals-14-00933" class="html-bibr">30</a>], Cong et al. [<a href="#B35-minerals-14-00933" class="html-bibr">35</a>], Cao et al. [<a href="#B39-minerals-14-00933" class="html-bibr">39</a>], Xie et al. [<a href="#B42-minerals-14-00933" class="html-bibr">42</a>], Chen et al. [<a href="#B44-minerals-14-00933" class="html-bibr">44</a>], Cong et al. [<a href="#B49-minerals-14-00933" class="html-bibr">49</a>], Li et al. [<a href="#B50-minerals-14-00933" class="html-bibr">50</a>], Chen et al. [<a href="#B51-minerals-14-00933" class="html-bibr">51</a>], Zhu et al. [<a href="#B52-minerals-14-00933" class="html-bibr">52</a>]).</p>
Full article ">Figure 2
<p>(<b>a</b>) Geological map of the Yingpanshan–Damanbie pluton with the Yingpanshan–Damanbie iREE deposit. (<b>b</b>) A profile of the regolith with iREE mineralization from the Yingpanshan–Damanbie iREE deposit.</p>
Full article ">Figure 3
<p>Characteristics of petrography and REE-bearing accessory minerals of syenogranite from the Yingpanshan–Damanbie iREE deposit in western Yunnan. (<b>a</b>) Photograph of a sample specimen; (<b>b</b>) photomicrograph; and (<b>c</b>) TIMA images of thin section; abbreviations: Kfs = K–feldspar, Qtz = quartz, Pl = plagioclase, Bt = biotite, Ep = Epidote.</p>
Full article ">Figure 4
<p>Characteristics of petrography and REE accessory minerals of monzogranite from the Yingpanshan–Damanbie iREE deposit in western Yunnan. (<b>a</b>) Photograph of a sample specimen; (<b>b</b>) photomicrograph; and (<b>c</b>) TIMA images of representative thin sections; abbreviations: Kfs = K–feldspar, Qtz = quartz, Pl = plagioclase, Bt = biotite, Ep = Epidote, Hb = hornblende.</p>
Full article ">Figure 5
<p>U–Pb concordia diagrams for (<b>a</b>) monzogranite (L–1–B6) and (<b>b</b>) syenogranite (L–1–B5) from the Yingpanshan–Damanbie iREE deposit and CL images of representative zircon grains.</p>
Full article ">Figure 6
<p>Plots of (<b>a</b>) (K<sub>2</sub>O + Na<sub>2</sub>O) versus SiO<sub>2</sub>, (<b>b</b>) A/NK versus A/CNK, (<b>c</b>) K<sub>2</sub>O versus SiO<sub>2</sub>, (<b>d</b>) K<sub>2</sub>O/Na<sub>2</sub>O versus SiO<sub>2</sub> of monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit.</p>
Full article ">Figure 7
<p>Plots of chondrite-normalized REE patterns (<b>a</b>,<b>c</b>) and primitive mantle (PM)-normalized spider diagrams (<b>b</b>,<b>d</b>) for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. Values for normalization are from Sun and McDonough [<a href="#B68-minerals-14-00933" class="html-bibr">68</a>], respectively. UCC = upper continental crust; LCC = lower continental crust; UCC and LCC data from Jahn et al. [<a href="#B69-minerals-14-00933" class="html-bibr">69</a>].</p>
Full article ">Figure 8
<p>Plots of (<b>a</b>) La/Sm versus La, (<b>b</b>) Zr/Hf versus SiO<sub>2</sub>, (<b>c</b>) FeO<sup>T</sup>/MgO versus (Zr + Y + Nb + Ce), (<b>d</b>) Zr versus 10,000 Ga/Al) [<a href="#B83-minerals-14-00933" class="html-bibr">83</a>], (<b>e</b>) Nb-Y-3Ga [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>]; and (<b>f</b>) Nb-Y-Ce [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>] for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. A = A-type granite; A<sub>1</sub> = A<sub>1</sub>-type granite; A<sub>2</sub> = A<sub>2</sub>-type granite; I = I-type granite; S = S-type granite; FG = Fractionated felsic granite; OGT = Unfractionated M-, I- and S-type granite.</p>
Full article ">Figure 8 Cont.
<p>Plots of (<b>a</b>) La/Sm versus La, (<b>b</b>) Zr/Hf versus SiO<sub>2</sub>, (<b>c</b>) FeO<sup>T</sup>/MgO versus (Zr + Y + Nb + Ce), (<b>d</b>) Zr versus 10,000 Ga/Al) [<a href="#B83-minerals-14-00933" class="html-bibr">83</a>], (<b>e</b>) Nb-Y-3Ga [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>]; and (<b>f</b>) Nb-Y-Ce [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>] for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. A = A-type granite; A<sub>1</sub> = A<sub>1</sub>-type granite; A<sub>2</sub> = A<sub>2</sub>-type granite; I = I-type granite; S = S-type granite; FG = Fractionated felsic granite; OGT = Unfractionated M-, I- and S-type granite.</p>
Full article ">Figure 9
<p>Plots of (<b>a</b>) Al<sub>2</sub>O<sub>3</sub>/(Fe<sub>2</sub>O<sub>3</sub><sup>T</sup> + MgO + TiO<sub>2</sub>) versus (Al<sub>2</sub>O<sub>3</sub> + Fe<sub>2</sub>O<sub>3</sub><sup>T</sup> + MgO + TiO<sub>2</sub>) (after Patiňo Douce. (1999) [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>]); (<b>b</b>) (Na<sub>2</sub>O + K<sub>2</sub>O) versus (FeO<sup>T</sup> + MgO + TiO<sub>2</sub>); (<b>c</b>) Mg<sup>#</sup> versus SiO<sub>2</sub>; and (<b>d</b>) ε<sub>Hf</sub>(t) versus U–Pb ages for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. (<b>b</b>) Compositional fields of experimental melts are from Patiño Douce [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>], Sylvester [<a href="#B85-minerals-14-00933" class="html-bibr">85</a>], Patiño Douce [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>], and Altherr et al. [<a href="#B86-minerals-14-00933" class="html-bibr">86</a>], respectively; (<b>c</b>) fields shown are as follows: pure crustal partial melts obtained in experimental studies by the dehydration melting of low-K basaltic rocks at 8–16 kbar and 1000–1050 °C [<a href="#B87-minerals-14-00933" class="html-bibr">87</a>]; pure crustal melts obtained in experimental studies by the moderately hydrous (1.7–2.3 wt.% H<sub>2</sub>O) melting of medium- to high-K basaltic rocks at 7 kbar and 825–950 °C [<a href="#B88-minerals-14-00933" class="html-bibr">88</a>]; mantle melts (basalts) and Quaternary volcanic rocks from the Andean southern volcanic zone [<a href="#B89-minerals-14-00933" class="html-bibr">89</a>]; melts from meta-igneous sources under crustal pressure and temperature conditions of 0.5–1.5 GPa and 800–1000 °C, respectively, which are based on the work completed by Wolf and Wyllie [<a href="#B90-minerals-14-00933" class="html-bibr">90</a>]; (<b>d</b>) data for the Gangdese belt from Ji et al. [<a href="#B91-minerals-14-00933" class="html-bibr">91</a>]; data for the southern Lhasa block from Jiang et al. [<a href="#B92-minerals-14-00933" class="html-bibr">92</a>], Ji et al. [<a href="#B93-minerals-14-00933" class="html-bibr">93</a>], Hou et al. [<a href="#B94-minerals-14-00933" class="html-bibr">94</a>], Zheng et al. [<a href="#B95-minerals-14-00933" class="html-bibr">95</a>], Zhu et al. [<a href="#B96-minerals-14-00933" class="html-bibr">96</a>], and Huang et al. [<a href="#B97-minerals-14-00933" class="html-bibr">97</a>]; data for the central Lhasa block from Hou et al. [<a href="#B94-minerals-14-00933" class="html-bibr">94</a>], Gao et al. [<a href="#B98-minerals-14-00933" class="html-bibr">98</a>], Zheng et al. [<a href="#B99-minerals-14-00933" class="html-bibr">99</a>], and Wang et al. [<a href="#B100-minerals-14-00933" class="html-bibr">100</a>]; data for the eastern Himalayan syntaxis from Chui et al. [<a href="#B101-minerals-14-00933" class="html-bibr">101</a>], Gou et al. [<a href="#B102-minerals-14-00933" class="html-bibr">102</a>], and Pan et al. [<a href="#B103-minerals-14-00933" class="html-bibr">103</a>]; and data for the Tengchong block including the Guyong area from Xu et al. [<a href="#B18-minerals-14-00933" class="html-bibr">18</a>], Xie et al. [<a href="#B42-minerals-14-00933" class="html-bibr">42</a>], Chen et al. [<a href="#B51-minerals-14-00933" class="html-bibr">51</a>], and Qi et al. [<a href="#B104-minerals-14-00933" class="html-bibr">104</a>].</p>
Full article ">
14 pages, 9837 KiB  
Article
Cenozoic Reactivation of the Penacova-Régua-Verin and Manteigas-Vilariça-Bragança Fault Systems (Iberian Peninsula): Implication in Their Seismogenic Potential
by Sandra González-Muñoz and Fidel Martín-González
Geosciences 2024, 14(9), 243; https://doi.org/10.3390/geosciences14090243 - 10 Sep 2024
Viewed by 449
Abstract
The Penacova-Régua-Verin (PRV) and the Manteigas-Vilariça-Bragança (MVB) are two of the longest faults of the Iberian Peninsula. These faults striking NNE–SSW, over lengths of >200 km, were developed during late-Variscan Orogeny and reactivated in response to the Alpine Cycle tectonics. Their tectonic evolution [...] Read more.
The Penacova-Régua-Verin (PRV) and the Manteigas-Vilariça-Bragança (MVB) are two of the longest faults of the Iberian Peninsula. These faults striking NNE–SSW, over lengths of >200 km, were developed during late-Variscan Orogeny and reactivated in response to the Alpine Cycle tectonics. Their tectonic evolution during Alpine compression (Cenozoic) and their implication in the active tectonic activity of Iberia are under discussion. Their recent tectonic activity is recorded in the vertical offset of geomorphological surfaces, in the associated pull-apart basins, and in M > 7 paleoseismic events. Based on the vertical surface offset of Pliocene surfaces (140–300 m for the MVB fault and 150–200 m for the PRV), together with the horizontal offset (1300–1600 m for MVBF fault and 600–1400 m for PRVF), we can conclude that they were reactivated as left-lateral strike-slip faults with a reverse component during the Pliocene (3.6 Ma)–present. These results indicate that these faults are not related to the strain transmission during the collision with Eurasia (Eocene–Oligocene). However, they are related to the intraplate strain of the southern collision with the African plate during the Upper Neogene. The estimated slip-rate is 0.2–0.5 mm/a for both faults. These slip-rates evidence important implications for the seismic hazard of this intraplate region. Full article
(This article belongs to the Section Structural Geology and Tectonics)
Show Figures

Figure 1

Figure 1
<p>Schematic geological map of the Iberian Massif and the location of the study area. Modified from [<a href="#B12-geosciences-14-00243" class="html-bibr">12</a>,<a href="#B29-geosciences-14-00243" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Tectonic models proposed for the PRV and MVB faults during the Cenozoic: (<b>a</b>) Model proposed by (e.g., [<a href="#B5-geosciences-14-00243" class="html-bibr">5</a>,<a href="#B9-geosciences-14-00243" class="html-bibr">9</a>]). (<b>b</b>) The model proposed by (e.g., [<a href="#B13-geosciences-14-00243" class="html-bibr">13</a>,<a href="#B14-geosciences-14-00243" class="html-bibr">14</a>]).</p>
Full article ">Figure 3
<p>(<b>a</b>) Panoramic view toward the south of the Vilariça pull-apart basin, showing the uplift of the western block across the fundamental surface. The number corresponds to the topographic profile in <a href="#geosciences-14-00243-f005" class="html-fig">Figure 5</a>. (<b>b</b>) Field picture of the pre-tectonic sediments (arkoses of Fm. Vilariça). (<b>c</b>) Field picture of the syn-tectonic sediments (conglomerates with blocks of quartzite and angular granites immersed in a sandy matrix, Fm. Bragança). (<b>d</b>) Field aspect of the fault breccia of the PRVF northern part. (<b>e</b>) Field picture of the Viana del Bollo basin and its syn-tectonic sediments, consisting mainly of sandy matrix conglomerates and quartzite cobbles. Note the 30° tilting, indicating the fault activity after its deposits. (<b>f</b>) Field picture of the syn-tectonic sediments in Viana do Bolo basin. Locations shown in <a href="#geosciences-14-00243-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Geological map of the traces and associated basins of the PRV and MVB faults. (<b>b</b>) Detailed map of the Chaves, Vila Real, and Telões basins. (<b>c</b>) Detailed map of the Mórtagua basin. (<b>d</b>) Detailed map of the Vilariça basin. (<b>e</b>) Detailed map of the Longroiva basin. (GLM) Galaico-Leoneses Mountains; (PCS) Portugal Central System.</p>
Full article ">Figure 5
<p>Main surfaces in the study area, with the vertical offset value measured and the localization of the profiles in <a href="#geosciences-14-00243-f006" class="html-fig">Figure 6</a>. Modified and improved from [<a href="#B21-geosciences-14-00243" class="html-bibr">21</a>,<a href="#B33-geosciences-14-00243" class="html-bibr">33</a>,<a href="#B34-geosciences-14-00243" class="html-bibr">34</a>].</p>
Full article ">Figure 6
<p>Profiles of surfaces through the PRV and MVB fault systems. The location of the profiles is shown in <a href="#geosciences-14-00243-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schema of the excess area technique. Modified from [<a href="#B45-geosciences-14-00243" class="html-bibr">45</a>]. (<b>b</b>) Digital elevation model map with the localization of the profiles (yellow lines) used for the area restoration technique. (<b>c</b>) Profiles and total uplift area for the Cantabrian Mountain (A-A′) and Galaico-Leoneses Mountains (B-B′).</p>
Full article ">
27 pages, 15384 KiB  
Article
Occurrence Mechanism and Controlling Factors of Shale Oil from the Paleogene Kongdian Formation in Cangdong Sag, Bohai Bay Basin, East China
by Binyu Ma, Qinhong Hu, Xiugang Pu, Shengyu Yang, Xuyang Wang, Wenzhong Han and Jiacheng Wen
J. Mar. Sci. Eng. 2024, 12(9), 1557; https://doi.org/10.3390/jmse12091557 - 5 Sep 2024
Viewed by 382
Abstract
Free oil, rather than adsorbed oil, is the main contributor to shale oil production with current development technologies, and assessing oil contents in different occurrence states (adsorbed oil vs. free oil) is a critical component in evaluating the economics of shale wells and [...] Read more.
Free oil, rather than adsorbed oil, is the main contributor to shale oil production with current development technologies, and assessing oil contents in different occurrence states (adsorbed oil vs. free oil) is a critical component in evaluating the economics of shale wells and plays. Although various methodologies have been developed, there are still some fundamental issues in assessing the oil contents in different occurrence states in shale. In this study, a new method was developed to estimate the adsorbed and free oil contents in the Second Member of the Eocene Kongdian Formation (Ek2) shales in Cangdong Sag, Bohai Bay Basin. This method combines the results of standard Rock-Eval pyrolysis and multi-step Rock-Eval pyrolysis with thin section petrography, X-ray diffraction for mineralogy, total organic carbon analyses, field emission scanning electron microscopy for pore morphology, and pore structure analyses by nitrogen physisorption and mercury intrusion porosimetry. Nine lithofacies were identified in a total of 50 shale samples, and the results show that the adsorbed and free oil are mainly contained in pores with diameters > 20 nm, and their contents are mainly controlled by organic matter abundance and thermal maturity of shales. While pore space volume influences the storage of shale oil, it is not a major determinant. Models of shale oil occurrence and its evolution are proposed, suggesting that the high S1 contents of organic-rich and -fair shales, which the latter resulted from oil migration, are the most favorable exploration targets of Ek2 shales. The findings of this study will help prioritize shale oil exploration targets in Ek2 shales. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Cangdong Sag and three sampling wells and stratigraphic column of Ek<sub>2</sub> in Cangdong Sag (modified from [<a href="#B27-jmse-12-01557" class="html-bibr">27</a>]).</p>
Full article ">Figure 2
<p>Ternary mineralogy classification of Ek<sub>2</sub> shales from three wells.</p>
Full article ">Figure 3
<p>Thin section petrography of typical Ek<sub>2</sub> shales: (<b>a</b>) Laminated structure, G108-8 well, 3183.08 m, plane polarized light; (<b>b</b>) Laminated structure, G108-8 well, 3183.08 m, crossed polarized light; (<b>c</b>) Laminated structure, G108-8 well, 3235.44 m, plane polarized light; (<b>d</b>) Laminated structure, G108-8 well, 3235.44 m, crossed polarized light; (<b>e</b>) Laminated structure, GD14 well, 4103.11 m, plane polarized light; (<b>f</b>) Laminated structure, GD14 well, 4103.11 m, crossed polarized light; (<b>g</b>) Massive structure, GD12 well, 3831.92 m, plane polarized light; (<b>h</b>) Massive structure, GD14 well, 4126.07 m, plane polarized light.</p>
Full article ">Figure 4
<p>Organic matter abundance of organic-rich shales related to lithofacies.</p>
Full article ">Figure 5
<p>SEM images of pore morphologies in Ek<sub>2</sub> shales: (<b>a</b>) G108-8 well, 3212.16 m; (<b>b</b>) GD12 well, 3894.93 m; (<b>c</b>) GD14 well, 4082.09 m; (<b>d</b>) G108-8 well, 2949.65 m; (<b>e</b>) G108-8 well, 3204.05 m; (<b>f</b>) G108-8 well, 3212.16 m; (<b>g</b>) G108-8 well, 3212.16 m; (<b>h</b>) GD14 well, 4082.09 m; (<b>i</b>) G108-8 well, 3050.26 m; (<b>j</b>) GD14 well, 4082.09 m; (<b>k</b>) GD14 well, 4136.21 m; (<b>l</b>) G108-8 well, 3204.05 m; (<b>m</b>) G108-8 well, 3204.05 m; (<b>n</b>) G108-8 well, 3212.16 m; (<b>o</b>) G108-8 well, 3212.16 m; (<b>p</b>) G108-8 well, 3212.16 m; (<b>q</b>) GD12 well, 3833.48 m; (<b>r</b>) GD14 well, 4095.27 m. Interp P: interparticle pore; Intrap P: intraparticle pore; Dis Intrap P: dissolved intraparticle pore; Dis Interp P: dissolved interparticle pore; OM P: organic matter pore; Micro F: micro-fracture; Q: quartz; F: feldspar; CAL: calcite; DOL: dolomite; Clay: clay minerals.</p>
Full article ">Figure 6
<p>(<b>a</b>) Nitrogen adsorption-desorption curves; and (<b>b</b>) Pore size distributions of typical organic solvent-extracted Ek<sub>2</sub> shales.</p>
Full article ">Figure 7
<p>Pore structure parameters of organic solvent-extracted shales from LNA analyses; (<b>a</b>) Specific surface area and pore volume; (<b>b</b>) Volume of different types of pores.</p>
Full article ">Figure 8
<p>Pore structure parameters of organic solvent-extracted shales from MIP analyses; (<b>a</b>) Porosity and total pore volume; (<b>b</b>) Volume of different types of pores.</p>
Full article ">Figure 9
<p>Comparison of the total oil content of shales of different lithofacies.</p>
Full article ">Figure 10
<p>Schematic diagram of pyrolysis spectra of as-received and solvent-extracted shales and division of shale oil into different occurrence states (FID: flame ionization detector).</p>
Full article ">Figure 11
<p>Comparison of the (<b>a</b>) adsorbed oil content and (<b>b</b>) free oil content of shales of different lithofacies.</p>
Full article ">Figure 12
<p>Correlations between pore volume increment and residual oil content: (<b>a</b>) Pore volume increment from LNA vs. residual oil content of organic-rich shales; (<b>b</b>) Pore volume increment from MIP vs. residual oil content of all shale samples.</p>
Full article ">Figure 13
<p>Relative proportions of residual oil in pores with different types.</p>
Full article ">Figure 14
<p>Correlations between total oil content and (<b>a</b>) TOC, (<b>b</b>) S<sub>1</sub>, and (<b>c</b>) total pore volume.</p>
Full article ">Figure 15
<p>Relationships between TOC, S<sub>1</sub>, S<sub>2</sub> and adsorbed oil and free oil contents. (<b>a</b>–<b>c</b>) The relationships between TOC, S<sub>1</sub>, S<sub>2</sub> and adsorbed oil content, respectively; (<b>d</b>–<b>f</b>) the relationships between TOC, S<sub>1</sub>, S<sub>2</sub> and free oil content, respectively.</p>
Full article ">Figure 16
<p>Relationships between TOC-normalized adsorbed oil content, TOC-normalized free oil content, and siliceous, calcareous, and clay minerals. (<b>a</b>,<b>c</b>,<b>e</b>) The relationships between siliceous, calcareous, clay minerals and TOC-normalized adsorbed oil content, respectively; (<b>b</b>,<b>d</b>,<b>f</b>) the relationships between siliceous, calcareous, clay minerals and TOC-normalized free oil content, respectively.</p>
Full article ">Figure 17
<p>Correlation between adsorbed oil content and parameters of pore structure. (<b>a</b>,<b>b</b>) The relationships between specific surface area, porosity and adsorbed oil content, respectively; (<b>c</b>–<b>e</b>) the relationships between non-seep-pore, potential seep-pore, seep-pore volumes and adsorbed oil content, respectively.</p>
Full article ">Figure 18
<p>Correlation between free oil content and parameters of pore structure. (<b>a</b>,<b>b</b>) The relationships between specific surface area, porosity and free oil content, respectively; (<b>c</b>–<b>e</b>) the relationships between non-seep-pore, potential seep-pore, seep-pore volumes and free oil content, respectively.</p>
Full article ">Figure 19
<p>Models of shale oil occurrence for (<b>a</b>) adsorbed and free oil distribution in shales with lower maturity and (<b>b</b>) adsorbed and free oil distribution in shales with higher maturity.</p>
Full article ">
22 pages, 4832 KiB  
Article
Cenozoic Carbon Dioxide: The 66 Ma Solution
by Patrick Frank
Geosciences 2024, 14(9), 238; https://doi.org/10.3390/geosciences14090238 - 3 Sep 2024
Viewed by 2184
Abstract
The trend in partial pressure of atmospheric CO2, P(CO2), across the 66 MYr of the Cenozoic requires elucidation and explanation. The Null Hypothesis sets sea surface temperature (SST) as the baseline driver for Cenozoic P(CO2). The crystallization [...] Read more.
The trend in partial pressure of atmospheric CO2, P(CO2), across the 66 MYr of the Cenozoic requires elucidation and explanation. The Null Hypothesis sets sea surface temperature (SST) as the baseline driver for Cenozoic P(CO2). The crystallization and cooling of flood basalt magmas is proposed to have heated the ocean, producing the Paleocene–Eocene Thermal Maximum (PETM). Heat of fusion and heat capacity were used to calculate flood basalt magmatic Joule heating of the ocean. Each 1 million km3 of oceanic flood basaltic magma liberates ~5.4 × 1024 J, able to heat the global ocean by ~0.97 °C. Henry’s Law for CO2 plus seawater (HS) was calculated using δ18O proxy-estimated Cenozoic SSTs. HS closely parallels Cenozoic SST and predicts the gas solute partition across the sea surface. The fractional change of Henry’s Law constants, HnHiHnH0 is proportional to ΔP(CO2)i, and HnHiHnH0×P(CO2)+P(CO2)min, where ΔP(CO2) = P(CO2)max − P(CO2)min, closely reconstructs the proxy estimate of Cenozoic P(CO2) and is most consistent with a 35 °C PETM ocean. Disparities are assigned to carbonate drawdown and organic carbon sedimentation. The Null Hypothesis recovers the glacial/interglacial P(CO2) over the VOSTOK 420 ka ice core record, including the rise to the Holocene. The success of the Null Hypothesis implies that P(CO2) has been a molecular spectator of the Cenozoic climate. A generalizing conclusion is that the notion of atmospheric CO2 as the predominant driver of Cenozoic global surface temperature should be set aside. Full article
(This article belongs to the Section Sedimentology, Stratigraphy and Palaeontology)
Show Figures

Figure 1

Figure 1
<p>Cenozoic times series plot of (<b>a</b>) (blue line), the δ<sup>18</sup>O proxy estimate of Cenozoic sea surface temperature, <span class="html-italic">T<sub>S</sub></span> [<a href="#B33-geosciences-14-00238" class="html-bibr">33</a>]; (<b>b</b>) (red line), calculated Henry’s Law constant for seawater, <span class="html-italic">H<sub>S</sub></span>, plotted with the ordinate decreasing upward. <span class="html-italic">H<sub>S</sub></span> varies with Cenozoic <span class="html-italic">T<sub>S</sub></span>.</p>
Full article ">Figure 2
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi mathvariant="normal">H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">H</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">H</mi> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, (blue points); <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mo>(</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <mo>)</mo> <mo>/</mo> <mi mathvariant="normal">x</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <mo>(</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <mo>)</mo> <mo>/</mo> <mi mathvariant="normal">x</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mo>(</mo> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <mo>)</mo> <mo>/</mo> <mi mathvariant="normal">x</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <mo>(</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <mo>)</mo> <mo>/</mo> <mi mathvariant="normal">x</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, (red line); <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, (black line); where P(CO<sub>2</sub>) is the historical atmospheric time series (ppm) since 1850. (<b>b</b>) (Blue line), Henry’s Law constant divided by the historical atmospheric CO<sub>2</sub> time series plotted against CO<sub>2</sub> (ppm). The structure reflects the impact of the natural variation of SST on the Henry’s Law constants (cf. h<sub>G</sub> in Equation (4)); (red line), linear least squares fit: H/P = 0.03779 − 8.466 × 10<sup>−6</sup> × CO<sub>2</sub> (ppm); r<sup>2</sup> = 0.84. (<b>c</b>) <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">P</mi> <mo>(</mo> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </msub> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, (red line), where P(CO<sub>2</sub>) is the historical record in ppm; (dashed red line), weighted fit through the data (see Sources and Methods Section); <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mo>/</mo> <mi mathvariant="normal">P</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> <mo>−</mo> <mo>(</mo> <mi mathvariant="normal">H</mi> <mo>/</mo> <mi mathvariant="normal">P</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> </mrow> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mo>/</mo> <mi mathvariant="normal">P</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mi mathvariant="normal">n</mi> </mrow> </msub> <mo>−</mo> <mo>(</mo> <mi mathvariant="normal">H</mi> <mo>/</mo> <mi mathvariant="normal">P</mi> <msub> <mrow> <mo>)</mo> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> (blue line), slope = 1.00, where P is total pressure in atm, including seawater vapor pressure plus P(CO<sub>2</sub>) (cf. Equation (7): y<sub>i</sub>/x<sub>i</sub> = H<sub>i</sub>/P).</p>
Full article ">Figure 3
<p>(<b>a</b>) (Blue points), P(CO<sub>2</sub>) (ppm) as calculated under the Null Hypothesis (Equation (14)), with <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msubsup> <mrow> <mi mathvariant="normal">P</mi> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> <mrow> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">l</mi> </mrow> </msubsup> </mrow> </semantics></math> = 862 ppm and the fractional ratio of the temperature-varying Henry’s Law constant for seawater (<a href="#geosciences-14-00238-f001" class="html-fig">Figure 1</a>). The red line is a 1% weighted Lowess smooth. Inset (<b>b</b>): expansion of the most recent 7.5 Ma.</p>
Full article ">Figure 4
<p>(<b>a</b>) (Red line), Henry’s Law estimate of the evolution of atmospheric P(CO<sub>2</sub>) with varying SST and [CO<sub>2</sub>]<sub>ocean</sub> held constant at the initial Cenozoic concentration of 3.41 × 10<sup>−5</sup> M; (blue line), Null Hypothesis estimate of the trend of P(CO<sub>2</sub>) constrained to the net proxy Cenozoic <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msubsup> <mrow> <mi mathvariant="normal">P</mi> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> <mrow> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">l</mi> </mrow> </msubsup> </mrow> </semantics></math> = 862 ppm. Inset (<b>b</b>): expansion of the first 25 Ma of the Cenozoic.</p>
Full article ">Figure 5
<p>(<b>a</b>) (Red line), Null Hypothesis estimate of the trend in atmospheric CO<sub>2</sub> during the Cenozoic; (black line), a 1% weighted Lowess smooth of the estimate; (blue points), Cenozoic P(CO<sub>2</sub>) as estimated using marine core alkenone and boron isotope proxies [<a href="#B2-geosciences-14-00238" class="html-bibr">2</a>]; and (blue line), a 5% weighted Lowess smooth of the proxy-derived trend in Cenozoic P(CO<sub>2</sub>). Inset (<b>b</b>): expansion of the most recent 7.5 Ma.</p>
Full article ">Figure 6
<p>(<b>a</b>) (Blue points), the Paleocene–Eocene proxy SST reconstruction of Bijl et al. (2009) [<a href="#B60-geosciences-14-00238" class="html-bibr">60</a>]; (red line), the Hansen et al. (2013) proxy SST reconstruction [<a href="#B33-geosciences-14-00238" class="html-bibr">33</a>]. The two reconstructions merge across 40 to 37 MYr BP. (<b>b</b>) (Pink points), proxy P(CO<sub>2</sub>) reconstruction of Rae et al. (2021), using marine core alkenone and boron isotope proxies [<a href="#B2-geosciences-14-00238" class="html-bibr">2</a>]; (purple line), 5% weighted Lowess fit through the points of the proxy; (blue line), estimate of Cenozoic P(CO<sub>2</sub>) (ppm) calculated using Equation (14), including the PETM SSTs of Bijl et al. [<a href="#B60-geosciences-14-00238" class="html-bibr">60</a>]; and (red line), piece-wise weighted Lowess fit to the Cenozoic CO<sub>2</sub> estimate. The weights were 15% over the reconstruction of Bijl et al., (65–37 MYr BP) [<a href="#B60-geosciences-14-00238" class="html-bibr">60</a>], and 1% over 37–0 MYr BP. Inset (<b>c</b>): expansion of the most recent 7.5 Ma.</p>
Full article ">Figure 7
<p>Double-y plot of the increase in P(CO<sub>2</sub>) during the MCO. (× points), proxy P(CO<sub>2</sub>) trend across the MCO [<a href="#B2-geosciences-14-00238" class="html-bibr">2</a>]; (circular points plus black line), the proxy P(CO<sub>2</sub>) trend and the Lowess smooth (cf. <a href="#geosciences-14-00238-f004" class="html-fig">Figure 4</a> and <a href="#geosciences-14-00238-f005" class="html-fig">Figure 5</a>); (red line), the Null Hypothesis estimate of SST-limited P(CO<sub>2</sub>) (cf. <a href="#geosciences-14-00238-f003" class="html-fig">Figure 3</a>); and (blue line), the Lowess smooth of the Null Hypothesis estimate. Both ordinates include a 550-ppm range but have been vertically offset so as to bring the two background slopes into coincidence.</p>
Full article ">Figure 8
<p>(<b>a</b>) Henry’s Law constants for seawater reflecting the adjusted SSTs (see text). Note the descending ordinate. (<b>b</b>) (Red line plus points), VOSTOK ice core P(CO<sub>2</sub>) record; and (blue line), P(CO<sub>2</sub>) calculated using [CO<sub>2</sub>]<sub>ocean</sub> = 9.980 × 10<sup>−6</sup> M and the adjusted <span class="html-italic">H<sub>S</sub></span> reflecting an 11 °C glacial/interglacial temperature range (Equation (17)). (<b>c</b>) Method comparison: (blue line), replication of the VOSTOK P(CO<sub>2</sub>) calculated using the adjusted Henry’s Law constants (Equation (17)); and (orange line), VOSTOK P(CO<sub>2</sub>) calculated according to the Null Hypothesis using the ratio of adjusted <span class="html-italic">H<sub>S</sub></span> values and the VOSTOK <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msubsup> <mrow> <mi mathvariant="normal">P</mi> </mrow> <mrow> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> <mrow> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">l</mi> </mrow> </msubsup> </mrow> </semantics></math> (Equation (18)). The two estimates express the assumption that the rises and falls of glacial/interglacial atmospheric P(CO<sub>2</sub>) reflect re-equilibrations across the sea surface driven by time-varying SST, alone.</p>
Full article ">
18 pages, 7517 KiB  
Article
Springs of the Arabian Desert: Hydrogeology and Hydrochemistry of Abu Jir Springs, Central Iraq
by John A. Webb, Jaafar Jotheri and Rod J. Fensham
Water 2024, 16(17), 2491; https://doi.org/10.3390/w16172491 - 2 Sep 2024
Viewed by 978
Abstract
The Arabian Desert is characterised by very low rainfall and high evaporation, yet over 210 springs are on its northeastern edge in central Iraq along the Abu Jir lineament, which represents the western depositional margin of a foreland basin infilled by the floodplain [...] Read more.
The Arabian Desert is characterised by very low rainfall and high evaporation, yet over 210 springs are on its northeastern edge in central Iraq along the Abu Jir lineament, which represents the western depositional margin of a foreland basin infilled by the floodplain sediments of the Tigris and Euphrates Rivers; there is little evidence of faulting. The springs discharge from gently east-dipping Paleocene–Eocene limestones, either where groundwater flowpaths intersect the ground surface or where groundwater flow is forced to the surface by confining aquitards. Calculated annual recharge to the aquifer system across the Arabian Desert plateau (130–500 million m3) is significant, largely due to rapid infiltration through karst dolines, such that karst porosity is the primary enabler of groundwater recharge. The recharge is enough to maintain flow at the Abu Jir springs, but active management of groundwater extraction for agriculture is required for their long-term sustainability. The hydrochemistry of the springs is determined by evaporation, rainfall composition (high SO4 concentrations are due to the dissolution of wind-blown gypsum in rainfall), and plant uptake of Ca and K (despite the sparse vegetation). Limestone dissolution has relatively little impact; many of the springs are undersaturated with respect to calcite and lack tufa/travertine deposits. The springs at Hit-Kubaysa contain tar and high levels of H2S that probably seeped upwards along subvertical faults from underlying oil reservoirs; this is the only location along the Abu Jir lineament where deep-seated faults penetrate to the surface. The presence of hydrocarbons reduces the Hit-Kubaysa spring water and converts the dissolved SO4 to H2S. Full article
Show Figures

Figure 1

Figure 1
<p>Location of springs (blue dots) along the Abu Jir lineament, and modern cities (squares). For location, see <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 2
<p>Topography of central Iraq showing the three major geomorphic sub-divisions, springs (white dots) located along the Abu Jir lineament, and the locations of the cross-sections in <a href="#water-16-02491-f005" class="html-fig">Figure 5</a>. For location, see <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 3
<p>Spring exposed in the bed of Sawa Lake when the lake receded (see <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> for location).</p>
Full article ">Figure 4
<p>Geology of the desert plateau west of Abu Jir lineament, after refs. [<a href="#B23-water-16-02491" class="html-bibr">23</a>,<a href="#B26-water-16-02491" class="html-bibr">26</a>,<a href="#B27-water-16-02491" class="html-bibr">27</a>], showing location of <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> and <a href="#water-16-02491-f002" class="html-fig">Figure 2</a> (box on main figure) and the stratigraphic profiles in <a href="#water-16-02491-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>Geological cross-sections along the Abu Jir lineament (for locations, see <a href="#water-16-02491-f002" class="html-fig">Figure 2</a> and <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>), showing the hydrogeology of the springs. Note the vertical exaggeration (x82); the actual westwards dip of the strata is &lt;1°. Stratigraphy derived from the outcrop distribution and bore logs on the following 1:250,000 geological maps: Karbala [<a href="#B33-water-16-02491" class="html-bibr">33</a>]; Al Najaf [<a href="#B34-water-16-02491" class="html-bibr">34</a>]; Baghdad [<a href="#B35-water-16-02491" class="html-bibr">35</a>]; Al Birreet [<a href="#B36-water-16-02491" class="html-bibr">36</a>]; Al Ramadi [<a href="#B37-water-16-02491" class="html-bibr">37</a>]; Shithatha [<a href="#B38-water-16-02491" class="html-bibr">38</a>].</p>
Full article ">Figure 6
<p>Piper plot of compositions of Abu Jir springs (data from) [<a href="#B8-water-16-02491" class="html-bibr">8</a>,<a href="#B10-water-16-02491" class="html-bibr">10</a>,<a href="#B11-water-16-02491" class="html-bibr">11</a>]; note that many available spring compositions, including the Shinafiyah springs, were incomplete as published and could not be plotted.</p>
Full article ">Figure 7
<p>Schoeller plot (standardised to Cl) comparing the median compositions of Abu Jir springs with desert plateau rainfall (see <a href="#water-16-02491-t001" class="html-table">Table 1</a> for data).</p>
Full article ">Figure 8
<p>Groundwater spring at Hit showing floating spongy bitumen (see <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> for location).</p>
Full article ">Figure 9
<p>Stable isotope data for some Abu Jir springs; groundwater data from nearby wells shown for comparison; data from refs. [<a href="#B7-water-16-02491" class="html-bibr">7</a>,<a href="#B48-water-16-02491" class="html-bibr">48</a>].</p>
Full article ">
20 pages, 8558 KiB  
Article
Notes on the Ecology and Distribution of Species of the Genera of Bondarzewiaceae (Russulales and Basidiomycota) with an Emphasis on Amylosporus
by Shah Hussain, Moza Al-Kharousi, Dua’a Al-Maqbali, Arwa A. Al-Owaisi, Rethinasamy Velazhahan, Abdullah M. Al-Sadi and Mohamed N. Al-Yahya’ei
J. Fungi 2024, 10(9), 625; https://doi.org/10.3390/jof10090625 - 1 Sep 2024
Viewed by 845
Abstract
The family Bondarzewiaceae is an important and diverse group of macrofungi associated with wood as white rotting fungi, and some species are forest tree pathogens. Currently, there are nine genera and approximately 89 species in the family, distributed in tropical, subtropical, and temperate [...] Read more.
The family Bondarzewiaceae is an important and diverse group of macrofungi associated with wood as white rotting fungi, and some species are forest tree pathogens. Currently, there are nine genera and approximately 89 species in the family, distributed in tropical, subtropical, and temperate climates. To address the phylogenetic relationships among the genera, a combined ITS-28S dataset was subjected to maximum likelihood (ML), Bayesian inference (BI), and time divergence analyses using the BEAST package. Both ML and BI analyses revealed two major clades, where one major clade consisted of Amylosporus, Stecchericium, and Wrightoporia austrosinensisa. The second major clade is composed of Bondarzewia, Heterobasidion, Gloiodon, Laurilia, Lauriliella, and Wrightoporia, indicating that these genera are phylogenetically similar. Wrightoporia austrosinensisa recovered outside of Wrightoporia, indicating that this species is phylogenetically different from the rest of the species of the genus. Similarly, time divergence analyses suggest that Bondarzewiaceae diversified around 114 million years ago (mya), possibly during the Early Cretaceous Epoch. The genus Amylosporus is well resolved within the family, with an estimated stem age of divergent around 62 mya, possibly during the Eocene Epoch. Further, the species of the genus are recovered in two sister clades. One sister clade consists of species with pileate basidiomata and generative hyphae with clamp connections, corresponding to the proposed section Amylosporus sect. Amylosporus. The other consists of species having resupinate basidiomata and generative hyphae without clamps, which is treated here as Amylosporus sect. Resupinati. We provided the key taxonomic characters, known distribution, number of species, and stem age of diversification of each section. Furthermore, we also described a new species, Amylosporus wadinaheezicus, from Oman, based on morphological characters of basidiomata and multigene sequence data of ITS, 28S, and Tef1-α. With pileate basidiomata and phylogenetic placement, the new species is classified under the proposed A. sect. Amylosporus. An identification key to the known species of Amylosporus is presented. Ecology and distribution of species of the genera in the family are discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the sampling sites, the Dhofar Governorate, located in south of Oman, the region consisting of lush wades and mountains, the photos showing some parts of the study area (Wadi Darbat with GPS coordinates: 17°07′ N, 54°43′ E; Wadi Jarzeez GPS coordinates: 17°13′ N, 54°05′ E; Wadi Rakhyut GPS coordinates: 16°80′ N, 53°42′ E). The pink dots represent sampling sites.</p>
Full article ">Figure 2
<p>Maximum likelihood phylogeny of family <span class="html-italic">Bondarzewiaceae</span> based on combined ITS-28 sequences; values above the nodes are maximum likelihood bootstrap (BT) and Bayesian posterior probabilities (PPs), with <span class="html-italic">Amylonotus labyrinthinus</span> (Yuan 1475) as outgroup. Species within the family were recovered in two clades. Clade-I consists of genus <span class="html-italic">Amylosporus</span>, <span class="html-italic">Stecchericium</span> and a species of <span class="html-italic">Wrightoporia W. austrosinens</span>, with moderate ML (BT 85%) and excellent BI support (PPs 1). Species in <span class="html-italic">Amylosporus</span> are further split into two sister clades: one sister clade with species having pileate basidiomata and clamped generative hyphae, representing the proposed section <span class="html-italic">A</span>. sect. <span class="html-italic">Amylosporus</span>; and the other sister clade with taxa having resupinate to effused-reflexed basidiomata and generative hyphae without clamp connections, making the section <span class="html-italic">A</span>. sect. <span class="html-italic">Resupinati</span>. Clade-II consisted of genera: <span class="html-italic">Bondarzewia</span>, <span class="html-italic">Heterobasidion</span>, <span class="html-italic">Gloiodon</span>, <span class="html-italic">Laurilia</span>, <span class="html-italic">Lauriliella</span>, and <span class="html-italic">Wrightoporia</span>, with excellent phylogenetic support in both analyses (BT 100%, PPs 1). The associated wood type, either angiosperm or gymnosperm are represented with their respective tree icon. Sequences of the new species <span class="html-italic">Amylosporus wadinaheezicus</span> are shown in bold font.</p>
Full article ">Figure 3
<p>Maximum clade credibility (MCC) tree of Bondarzewiaceae obtained from BEAST analysis based on ITS-28S dataset comprises 52 specimens including the outgroup taxon <span class="html-italic">Amylonotus labyrinthinus</span> (Yuan 1475). The dataset represents all the genera which belonging to Bondarzewiaceae according to Index Fungorum and Outlines of Fungi, except <span class="html-italic">Amylaria</span>, for which no sequence data are available. Taxa in the family recovered into two clades, in congruence to ML and BI analyses. One clade with an estimated stem age of diversification of approximately 94 mya, consisting of genus <span class="html-italic">Amylosporus</span>, <span class="html-italic">Stecchericium</span>, and <span class="html-italic">Wrightoporia austrosinens</span>. The horizontal dashed line separate <span class="html-italic">Amylosporus</span> for the rest of the genera of the family. The other clade with estimated stem age of divergence around 90 mya, consisting of <span class="html-italic">Bondarzewia</span>, <span class="html-italic">Heterobasidion</span>, <span class="html-italic">Gloiodon</span>, <span class="html-italic">Laurilia</span>, <span class="html-italic">Lauriliella</span>, and <span class="html-italic">Wrightoporia</span>. The 95% highest posterior density of divergence time estimations is marked by horizontal bars.</p>
Full article ">Figure 4
<p>Basidiomata of <span class="html-italic">Amylosporus wadinaheezicus</span>, (<b>A</b>,<b>B</b>); Mature basidiomata NHZ-22-004 (holotype), (<b>C</b>,<b>D</b>); HOD-23-012, (<b>E</b>); and DRB-23-009, (<b>F</b>–<b>H</b>). Pore surface under stereo microscope (NHZ-22-004).</p>
Full article ">Figure 5
<p>Illustration of anatomical characters of <span class="html-italic">Amylosporus wadinaheezicus</span> (holotype NHZ-22-004). (<b>A</b>) Basidia, (<b>B</b>) Basidiospores, (<b>C</b>) Contextual generative hyphae, (<b>D</b>) Gloeoplerous hyphae, (<b>E</b>) Contextaul skeletal hyphae, and (<b>F</b>) Tube generative hyphae. Scale bars: (<b>A</b>) = 10 µm, (<b>B</b>) = 5 µm, and (<b>C</b>–<b>F</b>) = 7 µm.</p>
Full article ">
14 pages, 27776 KiB  
Article
Coupling Relationship between Basin Evolution and Hydrocarbon Reservoirs in the Northern Central Myanmar Basin: Insights from Basin and Petroleum System Modeling
by Zengyuan Zhou, Wenxu Peng, Hefeng Sun, Kailong Feng and Weilin Zhu
J. Mar. Sci. Eng. 2024, 12(9), 1497; https://doi.org/10.3390/jmse12091497 - 29 Aug 2024
Viewed by 498
Abstract
The Myanmar region experienced the subduction of the Indian Ocean plate to the West Burma block and suffered from the land–land collision between the Indian continent and the West Burma block that occurred from the Late Cretaceous to the Cenozoic. Its tectonic evolution [...] Read more.
The Myanmar region experienced the subduction of the Indian Ocean plate to the West Burma block and suffered from the land–land collision between the Indian continent and the West Burma block that occurred from the Late Cretaceous to the Cenozoic. Its tectonic evolution has been complex; thus, oil and gas exploration is difficult, and the overall degree of research has been low. Recent exploration has been hindered by a lack of knowledge on the evolution of the petroleum system. To address this, we conducted hydrocarbon generation and accumulation modeling using both the 2D MOVE and Petro-Mod software 2017 for a complex tectonic section in the Northern Central Myanmar Basin. The results show that the maturity threshold depth of the Cretaceous source rocks in the study area is shallow, and the underground depth of 1200 m to 1400 m has reached the hydrocarbon generation threshold, indicating the start of hydrocarbon generation. Since 48 Ma, the Ro of the source rocks has reached 0.7%, became mature quite early. The Late Cretaceous Paleocene and Eocene formation, located in the southeastern part of the study area, migrated and accumulated hydrocarbons towards the western arc zone in the Eocene and Miocene, respectively. It is worth noting that although the oil and gas potential of each layer in the island arc uplift zone is relatively low, which is conducive to the migration and accumulation of oil and gas generated by the source rocks of the depression towards the island arc zone, shallow areas with developed extensional faults should be avoided. This study is the first to conduct a preliminary assessment and prediction of oil and gas resources, which will provide exploration guidance and reference for the study area and its surrounding areas in the future. Full article
(This article belongs to the Special Issue Exploration and Development of Marine Energy)
Show Figures

Figure 1

Figure 1
<p>Structural location, stratigraphic development characteristics of the CMB. (<b>a</b>) Geographic location of the CMB; (<b>b</b>) geological map of the CMB [<a href="#B23-jmse-12-01497" class="html-bibr">23</a>]; (<b>c</b>) comprehensive stratigraphic column of the study area.</p>
Full article ">Figure 2
<p>Seismic profile across the Northern Central Myanmar Basin. (<b>a</b>) The location of the section line A–B in the CMB is given in <a href="#jmse-12-01497-f001" class="html-fig">Figure 1</a>b; (<b>b</b>) the location of the section line C–D in the CMB is given in <a href="#jmse-12-01497-f001" class="html-fig">Figure 1</a>b; (<b>c</b>) the location of the section line E–F in the CMB is given in <a href="#jmse-12-01497-f001" class="html-fig">Figure 1</a>b.</p>
Full article ">Figure 3
<p>Distribution of formation residual thickness map in the Late Cretaceous to Miocene in the north part of the CMB. (<b>a</b>) Cretaceous–Paleocene (T<sub>80</sub>–T<sub>100</sub>); (<b>b</b>) Eocene (T<sub>70</sub>–T<sub>80</sub>); (<b>c</b>) Oligocene (T<sub>60</sub>–T<sub>70</sub>); (<b>d</b>) Miocene (T<sub>30</sub>–T<sub>60</sub>).</p>
Full article ">Figure 4
<p>Ro (%) distribution map of various source rocks in the Late Cretaceous to Eocene in the north part of the CMB. (<b>a</b>) During the Eocene Epoch; (<b>b</b>) during the Paleocene Epoch; (<b>c</b>) during the Late Cretaceous Epoch.</p>
Full article ">Figure 5
<p>Typical X1 well thermal evolution and burial history diagram in the north part of the CMB.</p>
Full article ">Figure 6
<p>Different degrees of hydrocarbon expulsion intensity of source rocks during the Late Cretaceous to Eocene in the north part of the CMB. (<b>a</b>,<b>b</b>) During the Late Cretaceous to Paleocene Epoch; (<b>c</b>) during the Eocene Epoch.</p>
Full article ">Figure 7
<p>Late Cretaceous–Paleocene and Eocene source rock expulsion history and cumulative expulsion curve in the north part of the CMB. (<b>a</b>) The source rock expulsion history during the Late Cretaceous to Paleocene Epoch; (<b>b</b>) the source rock expulsion history during the Eocene Epoch; (<b>c</b>) the source rock cumulative expulsion curve during the Late Cretaceous to Paleocene Epoch; (<b>d</b>) the source rock cumulative expulsion curve during the Eocene Epoch.</p>
Full article ">Figure 8
<p>Trends of hydrocarbon migration in different periods in the north part of the CMB. (<b>a</b>) During the Late Eocene Epoch; (<b>b</b>) during the Late Oligocene Epoch; (<b>c</b>) during the Late Miocene Epoch; (<b>d</b>) at present.</p>
Full article ">Figure 9
<p>Seismic data identification of multiple volcanic edifices in the north part of the CMB. (<b>a</b>) Seismic reflection characteristics of volcanic institutions; (<b>b</b>) tuned energy-inversion profile of Eocene formation near volcanic edifice.</p>
Full article ">Figure 10
<p>Restoration of balanced profile and reservoir migration model in the north part of the CMB. (<b>a</b>) During the Paleocene Epoch; (<b>b</b>) during the Eocene Epoch; (<b>c</b>) during the Oligocene Epoch; (<b>d</b>) at present.</p>
Full article ">
18 pages, 6982 KiB  
Article
Groundwater Quality Assessment at East El Minia Middle Eocene Carbonate Aquifer: Water Quality Index (WQI) and Health Risk Assessment (HRA)
by Abdel-Aziz A. Abdel-Aziz, Alaa Mostafa, Salman A. Salman, Ramadan S. A. Mohamed, Moustafa Gamal Snousy, Mohamed S. Ahmed, Mariacrocetta Sambito and Esam Ismail
Water 2024, 16(16), 2288; https://doi.org/10.3390/w16162288 - 14 Aug 2024
Viewed by 1276
Abstract
Around the world, groundwater supply is critical for vital needs such as drinking and irrigation. This work investigates groundwater in the carbonate aquifer of the Middle Miocene in the east El Minia area, Egypt. In this regard, thirty-two groundwater samples were collected. The [...] Read more.
Around the world, groundwater supply is critical for vital needs such as drinking and irrigation. This work investigates groundwater in the carbonate aquifer of the Middle Miocene in the east El Minia area, Egypt. In this regard, thirty-two groundwater samples were collected. The water samples were analyzed for Ca2+, Mg2+, Na+, K+, Cl, SO42−, NO3, CO2, HCO3, Fe, Mn, Cd, As, Cr, Cu, and Pb. Groundwater has been evaluated using two methods, which are water quality index (WQI) and health risk assessment (HRA). The predominant groundwater is soft water, and the samples range in salinity from fresh to slightly salty. The groundwater mostly falls into the alkaline water type. All the groundwater samples under study are deemed low quality for human consumption due to water contamination. Fe, Mn, Cd, Cu, and Pb have high HQnc values, which can result in non-carcinogenic health issues in adults, while Mn, Cu, and Pb can give rise to non-carcinogenic health issues in children. Full article
(This article belongs to the Special Issue Managing Water Resources Sustainably)
Show Figures

Figure 1

Figure 1
<p>Geological and geomorphological map and sampling sites of the area (after [<a href="#B27-water-16-02288" class="html-bibr">27</a>]).</p>
Full article ">Figure 2
<p>The climate of the area over the last 60 years, according to [<a href="#B29-water-16-02288" class="html-bibr">29</a>,<a href="#B30-water-16-02288" class="html-bibr">30</a>]; (<b>A</b>) is the climate from 1961 to 1990, while (<b>B</b>) is from 1991 to 2022.</p>
Full article ">Figure 3
<p>Geological cross-sections illustrate the typical faults that cut the studied sequence (<b>A</b>–<b>C</b>).</p>
Full article ">Figure 4
<p>Significations and nutrients concentrations in the studied wells.</p>
Full article ">Figure 5
<p>Ammonia (NH<sub>4</sub>) zoning map of the groundwater samples.</p>
Full article ">Figure 6
<p>Piper diagram for classification of the groundwater samples.</p>
Full article ">Figure 7
<p>Graphical projection of THI values in children and adults in the investigated area.</p>
Full article ">Figure 8
<p>Zonation map of (<b>A</b>) Total Hazard Index (THI) for adults; (<b>B</b>) arsenic non-carcinogenic health risks for adults; (<b>C</b>) chromium non-carcinogenic health risks for adults.</p>
Full article ">Figure 9
<p>Zonation map of (<b>A</b>) iron non-carcinogenic health risks for children; (<b>B</b>) cadmium non-carcinogenic health risks for children; (<b>C</b>) chromium non-carcinogenic health risks for children in the investigated area.</p>
Full article ">
21 pages, 34675 KiB  
Article
The Volcanic Rocks and Hydrocarbon Accumulation in the Offshore Indus Basin, Pakistan
by Jing Sun, Jie Liang, Jianming Gong, Jing Liao, Qingfang Zhao and Chen Zhao
J. Mar. Sci. Eng. 2024, 12(8), 1375; https://doi.org/10.3390/jmse12081375 - 12 Aug 2024
Viewed by 610
Abstract
To analyze the impact of volcanic rocks in the Offshore Indus Basin on hydrocarbon reservoir formation, seismic data interpretation, seismic data inversion, and sea–land correlation analysis were carried out. The results show that, longitudinally, volcanic rocks are mainly distributed at the top of [...] Read more.
To analyze the impact of volcanic rocks in the Offshore Indus Basin on hydrocarbon reservoir formation, seismic data interpretation, seismic data inversion, and sea–land correlation analysis were carried out. The results show that, longitudinally, volcanic rocks are mainly distributed at the top of the Cretaceous system or at the bottom of the Paleocene, and carbonate rock platforms or reefs of the Paleocene–Eocene are usually developed on them. On the plane, volcanic rocks are mainly distributed on the Saurashtra High in the southeastern part of the basin. In terms of thickness, the volcanic rocks revealed by drilling in Karachi nearshore are about 70 m thick. We conducted sparse spike inversion for acoustic impedance in the volcanic rock area. The results show that the thickness of the Deccan volcanic rocks in the study area is between 250 and 750 m which is thinning from southeast to northwest. Based on sea–land comparison and comprehensive research, the distribution of volcanic rocks in the Indian Fan Offshore Basin played a constructive role in the Mesozoic oil and gas accumulation in the Indus offshore. Therefore, in the Indian Fan Offshore Basin, attention should be paid to finding Mesozoic self-generated and self-stored hydrocarbon reservoirs and Cenozoic lower-generated and upper-stored hydrocarbon reservoirs. Full article
(This article belongs to the Section Geological Oceanography)
Show Figures

Figure 1

Figure 1
<p>Tectonic location of the Offshore Indus Basin (The blue box indicates the scope of study area. The red lines represent the boundary lines of the plates).</p>
Full article ">Figure 2
<p>The tectonic evolution of the Indian Plate (Reprinted with permission from Ref. [<a href="#B14-jmse-12-01375" class="html-bibr">14</a>]. Copyright 2012, Elsevier).</p>
Full article ">Figure 3
<p>Impedance histogram from Karachi South-1A.</p>
Full article ">Figure 4
<p>The distribution of Deccan volcanic rocks. (Reprinted with permission from Ref. [<a href="#B21-jmse-12-01375" class="html-bibr">21</a>]. Copyright 2005, Geological Society of America).</p>
Full article ">Figure 5
<p>Restoration map of the tectonics of the Indian plate and its surrounding areas 65 million years ago.</p>
Full article ">Figure 6
<p>The geological profile shows thin thickness of the Deccan volcanic rocks in the Lower Indus Basin (Reprinted with permission from Ref. [<a href="#B24-jmse-12-01375" class="html-bibr">24</a>]. Copyright 2017, Science Press).</p>
Full article ">Figure 7
<p>Well section of Deccan volcanic rocks encountered in the northeastern part of the Offshore Indus Basin. (Reprinted with permission from Ref. [<a href="#B7-jmse-12-01375" class="html-bibr">7</a>]. Copyright 2019, Elsevier.)</p>
Full article ">Figure 8
<p>Time-domain structural map of the top surface of the volcanic rock platform in the southeastern part of the Indian Fan Offshore Basin (modified according to reference [<a href="#B23-jmse-12-01375" class="html-bibr">23</a>]).</p>
Full article ">Figure 9
<p>Seismic profile across the volcanic platform in the southeastern of study area (see <a href="#jmse-12-01375-f001" class="html-fig">Figure 1</a> for the location of the profile).</p>
Full article ">Figure 10
<p>Inversion result of impedance for line 4 (see <a href="#jmse-12-01375-f001" class="html-fig">Figure 1</a> for the location of the profile).</p>
Full article ">Figure 11
<p>The thinness of volcanic rocks in the Offshore Indus Basin (see the blue dotted box in <a href="#jmse-12-01375-f001" class="html-fig">Figure 1</a> for the scope).</p>
Full article ">Figure 12
<p>The histogram of source–reservoir–caprock in the Lower Indus Basin (modified according to reference [<a href="#B32-jmse-12-01375" class="html-bibr">32</a>]).</p>
Full article ">Figure 13
<p>The distribution of oil and gas fields in the Lower Indus Basin in Pakistan’s land area (Data from IHS Markit).</p>
Full article ">Figure 14
<p>Analysis of carbonate reservoir in a well in the Offshore Indus Basin.</p>
Full article ">Figure 15
<p>Map of hydrocarbon accumulation events in the Offshore Indus Basin.</p>
Full article ">Figure 16
<p>Hydrocarbon accumulation model of the Offshore Indus Basin.</p>
Full article ">
13 pages, 10183 KiB  
Article
Tectonic Inversion and Deformation Differences in the Transition from Ionian Basin to Apulian Platform: The Example from Ionian Islands, Greece
by Avraam Zelilidis, Nicolina Bourli, Elena Zoumpouli and Angelos G. Maravelis
Geosciences 2024, 14(8), 203; https://doi.org/10.3390/geosciences14080203 - 31 Jul 2024
Viewed by 611
Abstract
The studied areas (the Ionian Islands: Paxoi, Lefkas, Kefalonia, and Zakynthos), are situated at the western ends of the Ionian Basin in contact with the Apulian Platform and named as Apulian Platform Margins. The proposed model is based on fieldwork, previously published data, [...] Read more.
The studied areas (the Ionian Islands: Paxoi, Lefkas, Kefalonia, and Zakynthos), are situated at the western ends of the Ionian Basin in contact with the Apulian Platform and named as Apulian Platform Margins. The proposed model is based on fieldwork, previously published data, and balanced geologic cross-sections. Late Jurassic to Early Eocene NNW–SSE extension, followed by Middle Eocene to Middle Miocene (NNW–SSE compression, characterizes the Ionian basin). The space availability, the distance of the Ionian Thrust from the Kefalonia transform fault and the altitude between the Apulian Platform and the Ionian Basin that was produced during the extensional regime were the main factors for the produced structures due to inversion tectonics. In Zakynthos Island, the space availability (far from the Kefalonia Transform Fault), and the reactivation of normal bounding faults formed an open geometry anticline (Vrachionas anticline) and a foreland basin (Kalamaki thrust foreland basin). In Kefalonia Island, the space from the Kefalonia Transform Fault was limited, and the tectonic inversion formed anticline geometries (Aenos Mountain), nappes (within the Aenos Mountain) and small foreland basins (Argostoli gulf), all within the margins. In Lefkas Island, the lack of space, very close to the Kefalonia Transform Fault, led to the movement of the Ionian Basin over the margins, attempting to overthrust the Apulian Platform. Because the obstacle between the basin and the platform was very large, the moving part of the Ionian Basin strongly deformed producing nappes and anticlines in the external part of the Ionian Basin, and a very narrow foreland basin (Ionian Thrust foreland basin). Full article
Show Figures

Figure 1

Figure 1
<p>Geological map of the studied area with the four studied cross-sections [<a href="#B28-geosciences-14-00203" class="html-bibr">28</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Geological map and (<b>b</b>) geomorphological map of Paxoi and Anti-Paxoi Island.</p>
Full article ">Figure 3
<p>(<b>a</b>) Geological map and (<b>b</b>) geomorphological map of Lefkas Island.</p>
Full article ">Figure 4
<p>(<b>a</b>) Geological map and (<b>b</b>) geomorphological map of Kefalonia Island.</p>
Full article ">Figure 5
<p>(<b>a</b>) Geological map and (<b>b</b>) geomorphological map of Zakynthos Island.</p>
Full article ">Figure 6
<p>Lithostratigraphic columns of (<b>a</b>) Ionian Basin and (<b>b</b>) Apulian Platform margins.</p>
Full article ">Figure 7
<p>Four cross-sections depict the major structures based on Google Earth Relief. For the abbreviations see the text and for the locations see <a href="#geosciences-14-00203-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 8
<p>Εvolutionary stages of development from rifting stage to present, applied to Paxoi and Anti-Paxoi Islands. (<b>a</b>) The rifting stage; (<b>b</b>) the change of the extensional to compressional regime with the reactivation of normal faults as reverse faults (inverted tectonic) and the gradual change from the Apulian platform margins to the forebulge area of the Ionian foreland and (<b>c</b>) the present morphology of Paxoi and Anti-Paxoi Islands with an open anticline geometry due to the Ionian thrust movement [<a href="#B37-geosciences-14-00203" class="html-bibr">37</a>].</p>
Full article ">Figure 9
<p>The block diagram illustrates how the tectonic inversion influenced the Apulian Platform Margins (APM) producing small, restricted foreland basins.</p>
Full article ">
Back to TopTop