Next Article in Journal
Artificial-Intelligence-Based Models Coupled with Correspondence Analysis Visualization on ART—Cases from Gombe State, Nigeria: A Comparative Study
Next Article in Special Issue
New Fossil Evidence Suggests That Angiosperms Flourished in the Middle Jurassic
Previous Article in Journal
Efficacy and Safety of Gel Immersion Endoscopic Mucosal Resection for Non-Pedunculated Colorectal Polyps
Previous Article in Special Issue
Unexpected Diversity of Xenoscelinae in Priabonian European Amber: The Third Xenosceline Species from Rovno Amber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Diagenetic and Biological Overprints in Geochemical Signatures of the Gigantoproductus Tertiary Layer (Brachiopoda): Assessing the Paleoclimatic Interpretation

by
José R. Mateos-Carralafuente
1,
Ismael Coronado
2,*,
Juncal A. Cruz
2,
Pedro Cózar
3,
Esperanza Fernández-Martínez
2 and
Sergio Rodríguez
1,3
1
Department of Geodynamics, Stratigraphy and Paleontology, Faculty of Geological Sciences, Complutense University of Madrid, c/ José Antonio Novais, 12, 28040 Madrid, Spain
2
Facultad de Ciencias Biológicas y Ambientales, Universidad de León, Campus Vegazana s/n, 24071 León, Spain
3
Instituto de Geociencias (CSIC, UCM), c/ Severo Ochoa 7, 28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
Life 2023, 13(3), 714; https://doi.org/10.3390/life13030714
Submission received: 10 February 2023 / Revised: 28 February 2023 / Accepted: 1 March 2023 / Published: 6 March 2023
(This article belongs to the Special Issue Recent Research on Palaeontology)

Abstract

:
Variations in the geochemical signatures of fossil brachiopod shells may be due to diagenesis and/or biological processes (i.e., ‘vital effects’). It is critical to characterise them in order to identify reliable shell areas suitable for paleoclimate studies. This investigation contributes to an in-depth understanding of geochemical variations in Gigantoproductus sp. shells (SW Spain, Serpukhovian age), throwing light onto the Late Paleozoic Ice Age interpretation. Microstructural, crystallographic, cathodoluminescence and geochemical (minor and trace elements, δ18O, δ13C, and strontium isotopes) characterisations have been performed on the tertiary layer of the ventral valve, to assess the preservation state. Poorly preserved areas exhibit microstructural and geochemical changes such as recrystallisation, fracturing and higher Mn and Fe enrichment. Moreover, these areas have a higher dispersion of ⁸⁶Sr, ⁸⁷Sr, δ18O and δ13C than well-preserved areas. Three structural regions have been identified in well-preserved areas of the ventral valve by differences in valve curvature and thickness, such as the umbonal and thick and thin regions. These regions have different proportions of Mg, S, Na, δ18O, and δ13C, which are interpreted as ‘vital effects’ and probably related to growth-rate differences during shell growth. The Gigantoproductus tertiary layer seems the most suitable for paleoclimate studies, because it retains the original microstructure and geochemical composition.

1. Introduction

The Late Paleozoic Ice Age (LPIA), elapsed from the Viséan to the Capitanian (~335 to 260 Ma) [1], was a dynamic climatic phenomenon that involved multiple glacial and interglacial events driven by the continental Pangea assembly, and low global atmospheric carbon dioxide concentrations (pCO2). These constraints, in turn, produced glacio eustatic variations, changes in ocean circulation, and finally, a turnover in marine flora and faunas [1,2,3,4,5,6,7,8]. The Mid-Carboniferous glacial peak represented an important change in seawater chemistry and ocean circulation [9] and led to one of the most important and selective crises in the Phanerozoic Era [10]. Low speciation and high extinction rates of marine faunas are the outstanding features during this period, particularly in the late Serpukhovian [11].
Most of the palaeoclimatological Carboniferous data are based on carbonate brachiopod δ18O and δ13C signatures, as paleotemperature and carbon-cycle proxies, respectively [9,12,13,14,15,16,17,18,19,20,21,22,23,24]. Rhynchonellid brachiopod shells are commonly selected for Carboniferous paleoclimate reconstructions, due to their worldwide distribution and abundance [13,19,20,21,25,26,27,28]. Carbonate brachiopod shells are composed of low-magnesium calcite (LMC), a polymorph which is relatively resistant to diagenesis [29].
Gigantoproductids (tribe Gigantoproductini [30]) are key, due to their great potential for palaeoclimatological studies because: (i) they were frequent during the Carboniferous [28,31,32]; (ii) they had global distribution [30]; (iii) they had large shell thickness and size [33], which allow the extraction of material for geochemical analyses more easily than with other fossil brachiopods [28,34,35]. Large shell size has been attributed to a possible chemosymbiotic lifestyle [35]. In spite of these advantages, few studies have focused on this fossil group for palaeoclimatological purposes [9,17,22,28,34]. These studies reported geochemical variations in gigantoproductid shells regarding other brachiopod genera, for example higher contents of Mg, S and Na than Martinia and Chorisitites [17,28]. Moreover, large fluctuations in stable isotopes (δ18O and δ13C) have been described in gigantoproductid shells concerning other brachiopod genera, such as Chorisitites, of the same age and from the same localities [28]. These variable compositions detected in gigantoproductuid shells raise several hypotheses about their origin: (i) they could be the result of diagenesis; (ii) they could be caused by kinetic and biological effects during shell growth; (iii) they could be result of a combination of both processes.
Despite the relative resistance of the brachiopod shells to diagenesis, it may obliterate the original microstructure and modify the shell geochemistry, making the palaeoclimatological interpretation difficult. In order to recognize diagenetic alteration, the identification of non-altered and altered areas in the brachiopod shell is needed. Altered regions are usually enriched or depleted in some trace elements such as Mn, Fe, Sr, and Na [36]; they exhibit isotope outliers; they show luminescence under cathodoluminescence [17], although non-luminescence is not always indicative of good preservation [37]; there is a lack of microstructures [36]; and the crystallographic orientation of biomineral structures is altered [38].
Despite the outstanding role of brachiopods shells in palaeoclimatological studies, some researchers have identified ‘vital effects’, considered herein as significant geochemical deviations from the thermodynamic equilibrium of seawater [39]. These variations have been recognized in extant brachiopods [40,41,42,43] and in well-preserved areas of fossil brachiopods (e.g., [22,24,34]). These geochemical variations in biominerals can be ascribed to: (i) biological processes [44] related to the metabolism and physiology of the organism (named ‘true vital effects’ by Pérez-Huerta and Andrus [45]); and (ii) bio-crystallization processes related to mechanisms of non-classical precipitation of biominerals [46,47]. In addition, both categories might also be affected by environmental factors such as temperature, salinity, and ocean acidification [46,48]. The selection among these hypotheses is complicated, and even an impossible process, due to the ‘vital effects’, which are currently poorly known mechanisms [45]. Geochemical variations derived from brachiopod shell growth are included within the category of biological processes such as kinetic effects or metabolic effects, including metabolic prioritization [24,40,42,49].
Growth rate and stable-isotopic fractionation covariation have been identified in δ18O and δ13C in extant and fossil brachiopods [24,35,40,49,50,51,52]. δ18O and δ13C variations between shell layers have been documented in extant and fossil brachiopods, and are considered as kinetic effects of the shell growth [24,43,49,50,51]. The tertiary layer is often selected for palaeoclimatological studies, due to its slower growth rate compared to the primary and secondary layers [9,28,34,36,43].
On the other hand, tailored δ26Mg fractionation has been detected between the primary and secondary layer in extant brachiopods [42]. Shell regions with higher growth rates are usually characterized as being 26Mg-enriched and 13C- and 18O-depleted [42]. Therefore, a covariation between growth rates and ion partitioning (e.g., minor and trace elements, MTE, such as Mg, Na, and S) has been identified in extant brachiopods [42], in which a heterogeneous Mg intra-shell distribution is described. Grossman et al. [36] described a covariation between Na and S from the inner to outer shells of fossil brachiopods, related to growth rates, with higher Na and S contents during higher growth rates. Moreover, a wide-ranging Mg variation in extant brachiopod shells has been documented [53,54,55]. The amount of MgCO3 in brachiopod shells varies ontogenetically during growth from Mg-enriched younger zones to progressively depleted older zones [55,56,57,58]. The Mg variation has a direct influence on the Mg/Ca ratio for the paleotemperature calculation, and needs further studies to establish robust knowledge of brachiopod growth rate and element covariation. This background highlights the need for a detailed characterization of the original microstructure and geochemistry of fossil brachiopod shells, prior to assessing the geochemical variation produced by diagenetic alteration and ‘vital effects’ [29,36].
The main aims of this study are: (i) to characterize the biogenic and diagenetic features of the tertiary layer of the Gigantoproductus sp. ventral valves (SW Spain, Serpukhovian Age), using a combination of structural, geochemical, and crystallographic techniques; (ii) to establish the most favorable zones for geochemical sampling, avoiding palaeoclimatological misinterpretations.

2. Geological Setting

The studied specimens were collected in the Guadiato Valley, coordinates 5°8′20″ W–38°14′ N, [59], a NW-SE elongated Carboniferous outcrop included in the Sierra Morena, at the southern Iberian Massif (Figure 1). Gigantoproductid samples were collected from the San Antonio section (San Antonio-La Juliana Unit), assigned to the Pendleian, Serpukhovian [60,61].
The Guadiato Valley (Figure 1) is an elongated zone where Carboniferous outcrops (late Viséan to early Westphalian) are separated by WNW–ESE faults [61]. It is divided into three tectono-sedimentary marine and paralic units: the Fresnedoso Unit, formed by siliciclastic rocks of the Viséan Age; the Sierra del Castillo Unit, formed mainly by carbonates of the Viséan Age, and the San Antonio-La Juliana Unit, formed by siliciclastics and carbonates of the Serpukhovian Age [59].
The San Antonio-La Juliana Unit has been interpreted in origin as a strike-slip basin with syntectonic sedimentation. It is composed of terrigenous and carbonate rocks and fossils of marine and paralic origin, which belong to several geographically close sedimentary environments. It comprises slope facies (hemipelagic sediments, olistoliths, debris flows and turbidites), platform facies (internal platform, tempestites with shallowing episodes), tidal-plain facies (intertidal plain and small lakes) and deltaic facies [60,62].
The San Antonio section, 146 m thick, is formed mainly by shales and siltstones with intercalated limestone and calcareous marlstones (Figure 1). The succession is open marine, with slope facies at the base that evolve into platform facies and deltaic facies in the upper part. The paleontological content is mostly represented by large-sized crinoids, gigantoproductids, bryozoans, rugose and tabulate corals, cyanoliths [62,63] and conodonts [61]. Based on foraminifera and conodonts, the Age is Pendleian (Early Serpukhovian; [61,64]).

3. Materials and Methods

3.1. Studied Material

Eleven specimens of Gigantoproductus sp. (Figure S1 in Supplementary Material) have been selected for this study, from the 2–4 horizon of the San Antonio section (Figure 1). The material (DMP-A301-1014-1 to DMP-A301-1014-11) is housed in the Paleontological Collections of the Paleontology Area (Complutense University of Madrid, UCM). The specimens were sectioned in two halves longitudinally, along the sagittal plane from the umbo to the commissure, when possible, and each slab was polished using sandpaper and 1 μm and 0.3 μm alumina. Polished samples were scanned in order to cross reference digitally the macroscopic features of the shells. One slab of each of the 11 specimens has been used for this study, and the other 11 slabs have been reserved for future studies. A total of twenty thin sections and five ultra-thin sections were prepared (sensu Coronado et al. [66]).

3.2. Microscopy Methods

The specimens were studied under a petrographic microscope, using scanning electron microscopy (SEM), computer-integrated polarization microscopy (CIP) and cathodoluminescence microscopy (CL), in order to select well-preserved areas for subsequent drilling.
A petrographic microscope, LM Leica DMLP, with a coupled camera Leica DC 300, was used for the purpose of characterizing the microstructure of the brachiopod shell.
Samples were extracted by natural breakage from the ventral valves of Gigantoproductus sp. between the muscle scars and the commissure, perpendicular and parallel to the valve growth. Moreover, a complete shell section was polished and etched in a 5% HCl solution for 20–25 s. These samples were coated with gold and analyzed by scanning electron microscopy (SEM), using a model JEOL JSM-820 working at 20 kV, located in the research facility of Geological Techniques of Complutense University of Madrid (UCM, Spain), in order to complete the microstructural characterization.
Thin sections were photographed under CL working at 16 kV and a current of 0.5 mA, in order to assess the intensity and distribution of luminescence. A cold cathodoluminescence probe model 8200 MK4 of Cambridge Image Technology Ltd. attached to a petrographic microscope model Eclipse E400 POL, with a camera, was used for this purpose. This microscope is located in the Geological Storage Division, Hydro-geochemical Group Ciemat (Madrid, Spain).
In order to evaluate shell preservation, the crystallographic organization of brachiopod shells was studied using computer-integrated polarization microscopy (CIP, [67]). This method for texture analysis and optical-orientation imaging has been applied in biomineralization studies of fossil bio-calcite with relevant crystallographic and structural results, highlighting it as a robust tool in diagenetic characterization [68,69,70]. This method determines the c-axis orientations of uniaxial minerals (such as calcite and quartz) from optical micrographs, displaying the results in pole figures and orientation images, using an RGB color-code that represents each orientation.
Seven CIP analyses in different samples and regions of the shell were performed, using an experimental petrographic Zeiss microscope with an automated rotation, and tilting system, made with the Arduino UNO microcontroller (open-source hardware) and controlled with software implemented in the LabVIEW environment (details in Coronado and Rodríguez [71]). The lowest ratio of pixel–µm reached with this setting in the micrographs is 1:0.05. Finally, the micrographs were processed using Image SXM software [72].
Crystal sizes were measured via micrographs using the plug-in ObjectJ 1.03w [73] of the free and open-source ImageJ 1.47v image-processing software [74].

3.3. Trace Elements and Isotope Analyses

Seven elements (Ca, Mg, Sr, S, Na, Mn and Fe) were analyzed using an electron microprobe analyzer (EMPA), with the analysis conducted on eleven thin sections (364 punctual analyses, 20 or 25 points per thin section), using a JEOL Superprobe, JXA-8900M with five wavelength-dispersive spectrometers, located in the Spanish National Centre for Electron Microscopy of the Complutense University of Madrid. Each point was analyzed with an accelerating voltage of 15 kV, a beam current of 10 nA and a spot size of 5 μm. Each analysis took an acquisition time of 45 s; because five elements can be analyzed simultaneously, the total time for analysis of the seven elements was 90 s. The following standards and detection limits were used in the EMPA: Ca: 201.5 ppm, Mg: 181.25 ppm (dolomite); Fe: 418.5 ppm, Mn: 439.5 ppm (siderite); Sr: 262.25 ppm (strontianite); Na: 167.75 ppm (albite) and S: 235.25 ppm (galena). In addition, the same seven elements were mapped. The EMPA mapping enables simultaneous analysis of different elements and the generation of distribution images for each element with 1 µm resolution. An accelerating voltage of 20 kV with a beam current of 100 nA and a spot size and step interval of 1 µm diameter (dwell time = 25 ms) were used.
Samples were powdered using a microdrill with an x-y micrometric plate attached to a binocular and a 500 μm tungsten carbide drill in selected areas, sampling for stable isotope analysis (δ18O, δ13C) and 87Sr/86Sr measurements. Stable isotope sampling requires only one drill hole to obtain enough material, whereas 87Sr/86Sr requires 30 to 40 drill holes.
Thermal ionization mass spectrometry (TIMS) was used in order to date the samples of study from the 87Sr/86Sr ratio. A total of 7 mg was obtained by the drilling of 4 samples of well-preserved zones and 2 samples of poorly preserved zones. The material was processed to obtain a Sr concentrate residue [75], which was loaded onto a Re filament by adding 1 μL of H3PO4 1M 4 and 2 μL of Ta2O5. The Sr isotopic ratios were analyzed on a TIMS-Sector 54® Mass Spectrometer, located in the Laboratory of Geochronology and Isotope Geochemistry (UCM), following a dynamic multicollection data-acquisition method for 10 blocks of 16 cycles each, with beam intensity mass of 88Sr at 3 V. The Sr analyses were corrected, to avoid possible interferences of 87Rb. The 87Sr/86Sr ratios were normalized with respect to the measured value of the ratio 86Sr/88Sr = 0.1194, in order to correct the possible fractionation of masses that the sample might have undergone during the filament loading and/or instrumental analysis. During the analysis of the samples, the isotopic standard of Sr (NBS 987) was analyzed repeatedly, and the following values were obtained: 0.710239 ± 1.7 × 10−5 (n = 8). These values were used to correct the analyses, taking into consideration the plausible drift referred to as the standard, and the certified value of the standard.
Stable isotope analyses (δ18O and δ13C) were undertaken using a triple-collector isotope-ratio mass spectrometer, Finnigan MAT 253 of the Stable Isotope Laboratory of the Department of Geological Sciences of the University of Michigan (USA). Forty-four samples were drilled, with 0.1 mg of each obtained, 30 from well-preserved zones and 14 from zones with evidence of poor preservation, previously characterized using optical microscopy, cathodoluminescence and CIP, and weighted and digested using H3PO4 at 77° ± 1 °C for 8 min in a Finnigan MAT Kiel IV. The analyses were calibrated using a better-fit regression line defined by two international standards, the NBS 18 (National Bureau of Standards; δ13C = −5.014‰ and δ18O = −23.2‰) and the 19 (National Bureau of Standards; δ13C = 1.95‰ and δ18O = −2.20‰). The data are given in ‰ notation, relative to the VPDB (Vienna Pee Dee Belemnite). The accuracy of the data was monitored through daily analyses, using a variety of carbonate-powder standards. At least 4 carbonate standards were reacted, analyzed and measured with an accuracy below ±0.1‰ for δ13C and δ18O.

4. Results

Collected specimens from the San Antonio section belong to horizons 2–4, which make them unsuitable for correlating geochemical variations with environmental factors (e.g., paleoseasonality). In order to differentiate diagenetic and biological overprints in Gigantoproductus shells, results have been divided into three main sections: shell microstructure, fossil preservation and geochemical characterization.

4.1. Shell Microstructure

Gigantoproductids have a concave-convex shell composed of a dorsal and ventral valve. The dorsal valve is thinner than the ventral valve. It has a mostly laminar microstructure, with lath crystals (sensu [76]), ~1–2 μm in width, grouped in laminae (sensu [34,76,77]) ~10 to 20 μm thick (Figure 2a); a few columnar layers, with smaller crystals than the ventral valve, appear within the laminar microstructure. The crystal size decreases towards the inner part of shell, as was recognized in the ventral valve.
The dorsal valve (Figure 2a) shows a similar thickness (~2–3 mm), which decreases towards the commissure, whereas the ventral valve has a noticeable thickness difference. The shell shape can be divided into three regions of the ventral valve based on the thickness differences: the umbonal region (U-region), the thick region (Tk-region) and the thin region (T-region) (Figure 2). The first region is more incurved than the rest of the shell, and is of ~8–10 mm in thickness, the second one is a thickened area, ~13–22 mm, following the U-region, and the third one is more planar than the U and the Tk region, with a ~5–7 mm thickness, which decreases from the thick region to the commissure.
In the ventral valve, two layers were distinguished: (i) the secondary or laminar layer and (ii) the tertiary or columnar layer. The laminar is formed by lath crystals, ~1 μm in width, arranged in packages of thin sheets ~10 to 40 μm in width (Figure 2f). The crystals are mostly tabular in appearance, with sharp contacts between the crystals (Figure 3e); the tertiary or columnar layer is formed of large columns (sensu [78]), with sizes ranging from 100 to 1700 μm in length and 40 to 150 μm in width (Figure 2e). Occasionally, the tertiary layer shows intercalation of laminar growth lines with a similar microstructure to the secondary layer embedded between the columns. The columnar crystals are composed of submicrometric steps stacked parallel to their growth direction (Figure 3d), forming terraces in the natural breakage (Figure 3g). Massive crystals growths in the columns occur commonly in the thick region (Figure 2e and Figure 3b).
Likewise, two types of microstructural changes parallel to the shell growth were identified in the tertiary layer, called growth lines: (i) a dashed and diffuse aspect (Figure 2(b2)) with several dark dots (i.e., opaque under transmitted light and in BSE images), passing through the columnar crystals, ~50 to 100 μm of thickness; (ii) a laminar microstructure (Figure 2(b1)), densely packed and with ~30 to 80 μm of thickness. These growth lines cut off the columnar crystals. A variation of the laminar microstructure with a granular appearance is observed in these growth lines.

4.2. Fossil-Shell Preservation

Fossil preservation has been addressed herein by the combination of cathodoluminescence, textural changes under both optical microscopy and SEM, the crystallographic arrangement of microstructures and the geochemical signatures (i.e., minor and trace elements and Sr isotopes). This set of techniques was applied to distinguish between well-preserved and poorly preserved shell regions, with the aim of comparing the luminescence and geochemical differences between the different shell regions and their relationship with the microstructure.

4.2.1. Textural Changes

Microtextural changes were recognized in some areas of the intra-shell surface, due to diagenetic processes. Fractures of the shell are usually oriented perpendicularly, or at a high angle to the shell surface, starting from the inner shell and moving to the outer shell, and are usually filled with micrite, sparite or iron oxides. On the other hand, thinner fractures, with micrite filling, parallel to the shell surface were observed in the center of some ventral valves. Some samples expose delamination processes near the shell surface.
Changes in the crystal size and the orientation of the tertiary layer are produced in adjacent zones of fractures or near to the edges of the shell. These areas contain a decreasing trend in crystal size towards the outer shell, with a mosaic appearance. Some areas exhibit micritization of the secondary layer and degrading neomorphism in the tertiary layer. These recrystallized zones are recognized by small and equigranular crystals, which totally obliterate the primary microstructure. Moreover, other zones contain partially dissolved crystals with a cloudy appearance, where the primary microstructure is roughly recognized. Rounded dark dots are opaque inclusions and are mostly concentrated in growth lines (Figure 2(b2)), whereas other scattered inclusions have sharp edges and different interference colors.

4.2.2. Crystallographic Arrangement

The observation of crystallographic patterns is a precise approach for the evaluation of diagenesis of polycrystalline skeletons formed by microcrystals, with preferred crystallographic orientations and a notable arrangement of supra-specialized structures [29]. This criterion has been previously applied to brachiopod shells, achieving excellent results in experimental diagenesis and in the evaluation of fossil shells [79,80,81]. In addition, crystallographic assessments of other calcium carbonate skeletons have been successful [65,67,68]. This approach is based on the principle that controlled mineralized shells have constrained crystallographic arrangements (i.e., preferred orientations), which, after diagenesis, can be compromised and substituted by random orientations and/or loss of textural features.
Seven areas of five ultra-thin sections of Gigantoproductus sp. ventral valves were studied after sampling for geochemical analysis (Figure 4). The CIP method is limited to assessing the crystallographic organization based on azimuth and inclination of the c-axis. Criteria to identify diagenetic alterations are based on abrupt changes in the c-axis orientation (highly disoriented), and textural changes with or without subtle c-axis variations.
The c-axis is strictly oriented perpendicular to the growth direction of Gigantoproductus shells (i.e., parallel to the radial axis of the valve), regardless of microstructure, shell region and texture, although some variations have been recognized. Columnar and laminar microstructures exhibit different c-axis orientations: parallel to the elongation axis of the crystals in the case of the columnar microcrystals, and perpendicular in the case of the laminar crystals (Figure 4 and Figure S2). In those areas where the laminar microstructure changes gradually to columnar, and vice versa (e.g., in a growth line), the c-axis is kept constant between the interconnected crystals (Figure 4b), similar to kinked crystals (sensu [82]).
All the studied areas, which include all the shell regions identified, demonstrate preferred orientations (based on pole figures), with quite narrow azimuthal dispersion. However, some distinctive arrangements have been observed in some shell regions. Usually, c-axis orientations in the tertiary layer show two narrow pole maxima, with an azimuthal dispersion between them of ca. 55° in the umbonal region (Figure S2e) and ca. 35° in the thick region (Figure S2h,k). On the other hand, thin regions with columnar and laminar microstructures exhibit a single pole maximum with a variable azimuthal dispersion: between 35° and 42° in the thin region (Figure 4b,e and Figure S2b).
Evidence of diagenetic alteration based on crystallographic orientation and textural changes occurs in some areas. In those areas, where the original microstructure is apparently well-preserved and lamellar twinning is notable, they are characterized by random c-axis orientations in twinned crystals and the surrounding areas (with hundreds of microns, Figure 4b,h). These areas correspond mostly to thin fractures without displacement, favored by microstructural changes (e.g., columnar to laminar or contact between columnar crystals, Figure 4h). In addition, in some of these areas the original microstructure has been completely obliterated by recrystallisation processes (Figure S2b), and these areas exhibit random azimuthal orientations, forming large patches with similar characteristics. On the other hand, in some areas the surrounding fractures are characterized by a degrading neomorphism with co-oriented microcrystals, compared to the main crystallographic orientation of the shell (Figure 4e), although the azimuthal dispersion of orientations in this region is larger than in individual crystals (i.e., columnar, and laminar crystals). In Figure 4g,h the diagenetic alteration is confined by a growth line, and involves only a part of the underlying crystals. This area is recrystallized because the original texture is lost, and random crystallographic orientation occurs.

4.2.3. Cathodoluminescence (CL)

Samples under CL exhibit almost exclusively non-luminescent areas (NL) (Figure 5), which correspond mostly to areas with well-ordered crystals of the tertiary layer (Figure 5a,c), some recrystallized zones (e.g., degrading neomorphism, Figure 5c), and fractures with iron-oxide cement (opaque under petrographic microscope, Figure 5b). Slightly luminescent areas (SL) are concentrated close to shell edges, in the laminar microstructure of the ventral and dorsal valves (Figure 5f,i), in the fracture zones with associated recrystallized areas (Figure 5d,e), in areas with degrading neomorphism (Figure 5i), in micritization zones close to the shell margin (Figure 5h), and in the growth lines (Figure 5g). High luminescent areas (L) correspond to the surrounding and infilling matrix (Figure 5f,h), the sparite cements of the inner space of the shell (Figure 5f,g), cemented borings (Figure 3j), large cemented fractures (Figure 5e,h), highly interconnected fractures or fracture zones (Figure 5g,h), and micritization areas close to the shell edge and some areas with neomorphism (Figure 5i).
The luminescence in the columnar crystals decreases from the edges to the inside (Figure 5), as well as in degrading neomorphism areas (Figure 5i). The luminescence of the growth lines increases in those areas affected by fractures, in comparison with the surrounding area (Figure 5d,g).

4.3. Geochemical Characterisation

Geochemical analyses, minor and trace-element (MTE) contents and isotope compositions, have been conducted in well- and poorly preserved areas, taking into consideration the luminescence under CL (Table 1, Figure 5), and the three shell regions identified (i.e., umbonal, thick and thin, Table 2, Table S2, and Table 3).

4.3.1. Major/Minor and Trace Elements (MTE)

MTE helps to evaluate the degree of geochemical alteration of brachiopods by comparing cathodoluminescence, crystallographic arrangement and petrographic microimages. The Gigantoproductus ventral valve shows different trace-element concentrations related to diagenetic alteration, layers, and shell regions (Figure S3, Table 1 and Table 2).
The proportion of trace elements varies with the function of the preservation state (Figure S3, Table 1). In general, well-preserved areas of the ventral valve show lower mean standard deviation of the Ca, Mg and S values than the poorly preserved areas. The NL well-preserved areas contain more Ca, less Mg and less dispersion of all elements, except Sr, than the NL poorly preserved areas. The SL well-preserved areas contain more Mn and Sr, less S, and exhibit lower mean standard deviation of Ca and Mg than the SL poorly preserved areas. The highest mean of Mn occurs in the luminescent areas (Figure S3, Table 1).
The mol% of MgCO3/CaCO3 of the NL and SL well-preserved areas, for each shell region, represents a linear correlation (r = −0.98 p(a) > 0.05, n = 249). Moreover, the average values of well-preserved areas is 98.02 mol% CaCO3 and 1.677 mol%MgCO3, distinctive of low-Mg calcite. In addition, differences between regions are recognized in the U-region (98.33 mol% CaCO3; 1.38 mol%MgCO3), Tk-region (98.48 mol% CaCO3; 1.26 mol%MgCO3), and the T-region (98.03 mol% CaCO3; 1.68 mol%MgCO3).
Statistics per element (Table 2) of the different brachiopod-shell regions demonstrate a higher average concentration of Ca and Mn in the umbonal region than in the rest of the regions. The thick region (Tk) exhibits less proportion of Mg and Mn than the U- and T-regions, but more Ca. The T-region contains slightly more Mg, Sr, and S than the U- and Tk-regions, but lower Na. Mg and Sr average values are higher in the secondary layer than in the tertiary layer, whereas Ca is lower in the secondary layer.
Ca, Mg, S and Na of all well-preserved areas are shown in boxplots in Figure 6. Mg is higher in the U- and T-regions, whereas Ca is higher in the Tk-region. Na is lower in the T-region; however, S is higher.
The Mg/Ca, S/Ca and Na/Ca transects measured in well-preserved areas of three individuals varies, from a higher concentration near the shell edge to a decreasing one towards the shell interior (Figure 7). In addition, the concentration of Mg, S and Na covaries in the measured transects. In poorly preserved areas (e.g., recrystallized areas) a decreasing-metal/Ca trend is not observed, and each element is not positively correlated (Figure 7c).

4.3.2. Isotope Geochemistry

The values of 87Sr/86Sr obtained from the TIMS range from 0.707830 to 0.707860 in the NL well-preserved areas (Table 3), while SL in the poorly preserved areas deviated from these values by ±0.00003. One sample is enriched in 87Sr, while another is depleted, in the cases of the well-preserved samples.
The δ13C–δ18O relationship can be subdivided according to the dispersion degree into poorly preserved and well-preserved areas and the consistency within the identified shell regions (Figure 7; Table 3). Dispersed values, with a variation of ~4.6‰ for δ18O and of ~3‰ for δ13C, occur in poorly preserved areas, whereas ~1.7‰ for the δ18O and ~1.3‰ δ13C occur in well-preserved areas. Poorly preserved areas show different alteration signatures: some samples have similar δ13C but are depleted in δ18O, others are depleted in δ13C and δ18O, and others are δ13C- and δ18O-enriched (Figure 8; Table 3).
A correlation with each identified shell region is recognized in well-preserved areas (i.e., the same drilled regions contain rather similar values; Figure S5). Thus, the δ18O values of well-preserved areas range from −3.8 to −2‰ and from 2.1 to 3.4‰ for δ13C. The U-region is where the data shows the lowest δ18O, with averages of −3.4‰ for δ18O and 2.6‰ for δ13C (Table 3). The Tk-region exhibits similar δ13C data to that of the U-region, but lower δ18O, with averages of −2.6‰ for δ18O and 2.6‰ for δ13C. Higher δ13C values are recognized in the T-region compared to the U-region, with averages of −3.2‰ for δ13C and 2.7‰ for δ18O.

5. Discussion

In this section, the preservation of the tertiary layer of Gigantoproductus ventral valves and the influence of diagenesis on the geochemical signatures are discussed. Moreover, the kinetic and biological effects that overprint the geochemical signatures suitable for paleoclimatic purposes, are assessed.

5.1. Fossil Preservation

The Gigantoproductus ventral valve can retain its original structural features at macro- and at microscale. The secondary and tertiary layers are well-defined by their microstructure, with a controlled preferred-crystallographic arrangement (the c-axis perpendicular to the shell edge). The microstructure is interpreted as unaltered because it retains its original morphological features, similar to those described in some extant brachiopods [82,83]. Likewise, submicrometric laminae have been recognized (Figure 3d) in columnar crystals of the tertiary layer, such as those described by Schmahl et al. [83] in the columnar crystals of Terebratulina septentrionalis. Furthermore, preserved nanocrystals, with a granular texture, were identified in some columnar microcrystals from natural breakage (Figure 3f,h). This finding suggests good preservation (i.e., the preservation of biogenic features) in fossil carbonate biominerals [29,84], because the nanocrystals are also characteristic of extant brachiopods [85].
Nonetheless, some areas show diagenetic alteration such as fractures, delamination, borings, micritization, and degrading neomorphism, with subsequent changes in the original crystallographic arrangement (Figure 4, Figure 5 and Figure S2).
The secondary and tertiary layers in Gigantoproductus shells are characterized by different microstructure and geochemical compositions, as was recognized in other extant and fossil brachiopods [36,43,86]. The secondary layer, in the Gigantoproductus of this study, is characterized by a laminar microstructure, with the elongate axis of crystals oriented parallel to the shell surface and with a small crystal size in comparison with the columnar microstructure of the tertiary layer, where the elongation and c-axis are oriented perpendicular to the shell surface and with larger crystals. Therefore, the morphological features of the secondary layer make it more prone to fracture, delamination, and recrystallisation. Within the tertiary layer, the thick region (Tk) of the valve is less affected by recrystallisation and fractures than the thin and umbonal regions (T, U). This may be due to their large crystal size and/or central shell position. On the other hand, the thin dorsal valve contains more fractures than the ventral valve, which makes it more susceptible to alteration and to more micritized patches (Table 4).
The secondary layer is highlighted by more intense luminescence under CL than the tertiary layer of the ventral valve (Figure 3). Similar differences in CL-luminescence intensity between the secondary and the tertiary layers had been recognized in Composita subtilita and Neospirifer pattersoni [87]. Although classically Mn and Fe are indicative of diagenetic alteration in calcite [88,89,90,91], not all luminescence, or its absence, can be correlated with diagenesis [65,78,79,92]. Biogenic carbonates can incorporate small amounts of Mn and Fe during their metabolic activity because of physicochemical variations of the environment in which they develop [92,93,94].
The luminescence differences in Gigantoproductus might be related to variations in the chemical composition; thus, the secondary layer is slightly enriched in Mn and Mg (Table 2), possibly related to more organic-matrix remains (Figure 3h). In addition, some fractures with a higher amount of Fe are non-luminescent under CL, and the growth lines with an apparently higher Mn concentration are luminescent under CL. This is partly due to the fact that Mn is a luminescence activator and Fe is an inhibitor of CL [91]. High Sr and Fe concentrations in some areas of the secondary layer suggest alteration by fluids favored by delaminated areas.
Slightly luminescent areas with a well-constrained crystallographic arrangement (considered here as well-preserved) have been identified in Gigantoproductus. These areas exhibit a luminescence pattern under CL that corresponds to the growth lines, and, in turn, show a slightly higher Mn content. This luminescence pattern under CL seems to be the result of organic matter and/or Mn [95] incorporated during shell growth. These areas also contain a higher Mg and S content, but no extensive Mn changes. The determination of small amounts of organic inclusions in growth lines (Figure 3h), and higher concentrations of Mg and Mn across the growth lines, suggest a combination of both sources, probably induced during the biocrystallization process. Similar luminescence patterns under CL were reported in the growth lines of the extant brachiopod Megerlia truncata [78], in fossil brachiopods [92,95] and in belemnite rostra [96]. Higher organic concentrations are recognized when the secretory regime of calcite is reduced [92]. This fact has been interpreted as growth cessation during environmental-stress events, produced by mantle anaerobiosis or by the acidification of calcification fluid by the closure of the valve [95,97,98]. Evaluation of these hypotheses deserves further study.
As described previously, not all non-luminescence areas of Gigantoproductus shells correspond to well-preserved zones. Crystals with disordered c-axis orientations have been identified in non-luminescent areas under petrographic microscopy. This may be due to slow recrystallisation processes, which keep the original microstructure but change the orientation of the c-axis. Examples of this process can be observed in coral skeletons [67,70]. Therefore, luminescent areas are not always indicative of diagenetic alteration [92] and non-luminescent areas are not always indicative of good preservation [29,37]. Banner and Kaufman [18] and Barbin and Gaspard [92] documented non-luminescent areas with altered 87Sr/86Sr and δ18O values in fossil brachiopod shells and slightly luminescent areas without evidence of alteration in an extant brachiopod.
The combination of CL and geochemical analyses helps to evaluate diagenetic alteration [65], but it is necessary to fully understand the original brachiopod-shell chemistry before using geochemical features such as a diagenetic indicator, in order to consider those biotic and abiotic factors that could control the original shell chemistry [36]. The chemical composition of the Gigantoproductus shells in this study is equivalent to that in the gigantoproductid reported by Popp et al. [17], Bruckschen et al. [20], Armendáriz et al. [9], Angiolini et al. [34,35] and Nolan [32]. The luminescence pattern under CL in well-preserved areas is similar to those described by Angiolini et al. [34,35]; additionally, a similar luminescence of diagenetically altered areas was described by Armendáriz et al. [9] and Nolan [32].
Minor and trace elements have been traditionally used to evaluate brachiopod-shell preservation by comparison between individuals and genera and with extant, unaltered brachiopod shells [17,28,36]. The NL and SL poorly preserved areas of the tertiary layer of the Gigantoproductus (Figure S3, Table 1) have ~800 ppm more Mg than the well-preserved areas. Moreover, the NL poorly preserved areas have ~800 ppm more Fe and ~120 ppm more Mn than the equivalent well-preserved areas. These values for the poorly preserved areas probably reflect the influence of Mg-, Fe-, and Mn-rich fluids during diagenesis in different stages (burial and meteoric waters). These data agree with the observations made by Grossman et al. [36] regarding altered brachiopod fossil shells, which are Fe- and Mn-enriched.
Additionally, the diagenetic alteration of valves modifies the isotopic record of δ18O, δ13C and 87Sr/86Sr. Poorly preserved areas exhibit δ18O and δ13C with larger standard deviations than those of the well-preserved areas (Table 3). Five samples of poorly preserved areas (Figure 8) are depleted in δ18O, with a similar depletion for δ13C; two samples are depleted in δ18O and δ13C; three samples are enriched in δ18O and δ13C; and two have similar amounts of δ18O, but show a δ13C enrichment with respect to the well-preserved areas.
In relation to Sr isotopes, well-preserved areas contain homogenous 87Sr/86Sr ratios, in contrast with poorly preserved areas (with one 87Sr/86Sr-enriched sample and one 87Sr/86Sr-depleted sample). Sr-isotope values from the Gigantoproductus (from 0.707830 to 0.707860) are equivalent to those shown by Bruckschen et al. (1999) in brachiopods from the Pendleian substage (mostly gigantoproductids shells, from 0.707828 to 0.707879), which validate the good preservation of the samples, except for two samples.
Different signatures of stable isotopes (δ18O and δ13C), trace elements (Mg, Mn, and Fe), and Sr isotopes suggest the co-existence of different diagenetic processes and stages acting on the Gigantoproductus shells, a fact which is also supported by the showcasing of microstructural and crystallographic changes.
On the other hand, the δ18O and δ13C trends in well-preserved areas can be explained by other mechanisms, such as kinetic and biological variations in ionic and isotopic values from the equilibrium of seawater (‘vital effects’).

5.2. Biological Overprint of Geochemical Signatures

‘Vital effects’ are geochemical deviations from the thermodynamic equilibrium of seawater produced by the vital processes of organisms, such as kinetic effects during ontogeny. Curry [99] observed three growth phases with different growth rates in Terebratulina retusa: a fast growth rate during the first stages of growth (in the umbonal region), followed by a constant growth period of three years and a progressively decreasing growth rate with brachiopod ageing. Variations in Mg across the shell have been widely studied in the literature as indicative of growth rates. Buening and Carlson [56] observed a higher Mg amount near the umbo of extant brachiopods, which decreases during brachiopod ageing.
The tertiary layer of the Gigantoproductus ventral valve can be divided into structural regions (umbonal, thick and thin regions) and growth stages (younger and older), because during ageing the shell grows in length (from umbo to commissure) and thickens (towards the interior of the valve), as a coupled process. Therefore, the U-, Tk- and T-regions must include younger (the outermost parts of the shell) and older (the innermost parts of the shell) areas, which may disguise some kinetic fractionation, considering the mean values of each area. For instance, structural regions exhibit differences in averaged Mg: the umbonal region has more Mg than the thick region (Tk), and the thin region (T) has the highest amount of Mg. On the other hand, transects analyzed for the function of distance exhibit a Mg depletion towards the interior from the outermost part of the ventral valve. Tk-region shows an exponential variation of Mg in contrast to T-region, which varies linearly.
Different Mg trends have been identified in extant brachiopod shells: an Mg-enrichment with ageing in the extant brachiopod shell Magellania venosa and a parabolic Mg trend with ageing in the extant brachiopod shells Liothyrella neozelanica and Gryphus vitreus [51]. Variations in Mg concentration of brachiopod shells seems to be species-specific [54,100], with different distribution trends among species [51] or within the same species [29,54]. Moreover, Rollion-Bard et al. [43] showed differences in the Mg/Ca incorporation across the shell in different extant brachiopod species, emphasizing the fact that Mg does not incorporate by a simple pathway during ontogeny. These differences may be related to different growth rates [56], crystallographic features [45], different proportions of organic components in the shell, seawater temperature and pH variations [42], and/or Mg exclusion from the calcification fluid [43]. Besides the Mg variation, S and Na vary from the outer edge of the ventral valve of Gigantoproductus towards the interior. The microstructure of the tertiary layer shows abundant smaller crystals located in the interphase between the secondary and tertiary layers (the initial biocrystallization event of the tertiary layer, Figure 2d), where nanograins and organic inclusions are frequent. Remarkably, these areas with small crystals are Mg- S- and Na-enriched, and during ageing the shell crystals enlarge and these elements decrease towards the interior.
Na in Gigantoproductus seems positively correlated with Mg and S (Figure 7) in the well-preserved areas. Na concentration in carbonates can be influenced by different processes that might explain this trend. For example, Na incorporation in abiogenic calcite and biogenic calcite may be favored by crystalline defects and by higher precipitation rates [101,102]. Moreover, Na can occupy an interstitial position in the carbonate-crystal lattice or can form part of the carbonate lattice by an altervalent substitution, creating a vacancy in the lattice [103,104]. On the other hand, Na can form part of the organic matter in biogenic carbonates; for instance, Rollion-Bard and Blamart [102] described a positive correlation between Na/Ca and Mg/Ca with the organic matter in bioaragonite in corals. Finally, a composition of carbonates can be strongly influenced by diagenesis, more than by environmental processes [105].
Mii et al. [106] showed a positive correlation between Mg, S and Na in fossil samples of Neospirifer dunbari, similar to the Gigantoproductus of this study, supporting the hypothesis that these trends are original signatures, indicative of well-preserved areas in fossils.
Mg, Na and S incorporation into a crystal lattice have been associated with crystal growth rates in crystal-growth experiments [101,107]. Moreover, growth rates associated with trace-element partitioning have been observed in both extant and fossil brachiopod shells [36,56,106]. Mii et al. [106] and Grossman et al. [36] related the high concentrations of these elements to seasonal growth rates, increasing during summer and decreasing during winter. Intra-shell small variations in Mg, Na and S in Gigantoproductus-shell transects (Figure 6) may be explained by this interpretation, despite the overprinted depletion trend from the shell edge towards the interior caused by brachiopod ageing.
All specimens of Gigantoproductus exhibit a depletion of δ13C and δ18O from older regions (the Tk- and T-regions) to younger regions (the U-region). It is remarkable that data from the thick region (Tk) are split into two clusters (Figure 8): one near the shell edge (younger), more depleted in δ13C, and another close to the shell interior (older), enriched in δ13C. Fractionation of δ13C and δ18O in Gigantoproductus shells exhibits an inverse trend to that observed in Mg, S and Na partitioning. Younger regions with higher growth rates are Mg-, S-, and Na-enriched and δ13C- and δ18O-depleted, whereas older zones are Mg-, S-, and Na-impoverished and δ13C- and δ18O-enriched. Preferential isotope fractionation has been previously described in extant and fossil brachiopods: for instance, Auclair et al. [49] described δ13C and δ18O depletion from the outermost (younger) to the innermost (older) part in the secondary layer of Terebratalia transversa. Similarly, Batt et al. [24] recognized δ13C depletion in the secondary layer of a fossil productid and δ13C and δ18O depletion in the tertiary layer from the outermost (younger) to the innermost (older) part of the genus Composita shell. Isotope fractionation was related to growth rate in brachiopod shells [24,49,50].
δ13C and δ18O direct- and inverse-covariations trends have been reported in extant and fossil brachiopod shells [24,42]. Moreover, the depletion trend of high growth-rate areas is not only linear, due to the parabolic trend, and inverse parabolic trends have been described in other brachiopod shells [51]. These variations in the partitioning and fractionation of ions and isotopes observed in different species and genera are called ‘vital effects’ and have been related to specific growth rates or metabolic changes in the brachiopods. Some authors were concerned about their influence on the shell geochemistry (e.g., Mg/Ca, δ13C and δ18O) and the possible paleoclimatic misinterpretations [50]. According to McConnaughey [108], ‘vital effects’ represent a noteworthy problem for isotopic geochemistry and paleoclimatology studies. The current results on Gigantoproductus geochemical variation underline the fact that they can be extended to the fossil record and, more specifically, to the interpretation of the Late Paleozoic Ice Age (LPIA). This requires the identification of ‘ideal’ sampling regions, in order to avoid differences in ion partitioning and isotope fractionation related to the differences in the growth rates (kinetic effects).
The Gigantoproductus regions less affected by “vital effects” (the sensu “equilibrium zone” or “plateau zone” identified by Perez-Huerta et al. [55] and Rollion-Bard et al. [43]) are the T-zone and the inner zone of the Tk-regions (Figure 7 and Figure 8), because of the low dispersion of the data. The umbonal region and external parts of the Tk-region of the Gigantoproductus ventral valves should be discarded, to avoid paleoclimatic misinterpretations (up to ~1.05‰ of δ18O and 14 mmol/mol of Mg/Ca). Herein, the Gigantoproductus tertiary layer offers great potential for paleoclimate studies, due to (i) a higher volume of shell substance compared with the secondary layer; (ii) growth-rate variability, which allows the characterization of the ‘vital effects’ of the shell; (iii) being less prone to diagenetic alteration than the secondary layer; and (iv) higher growth-line spacing, which allows for the avoidance of growth lines.

6. Conclusions

A comprehensive study of the microstructure, crystallography, and geochemistry (trace elements, δ18O, δ13C and 87Sr/86Sr) of the tertiary layer of the Gigantoproductus sp. ventral valves, has allowed the differentiation of areas with signs of diagenesis, from the primary biogenic features. In addition to fossil preservation, a special emphasis has been given to the characterization of the geochemical ‘vital effects’.
The Gigantoproductus shells consists of three layers, but only two are still preserved: the secondary layer with a laminar microstructure, and the tertiary layer with a characteristic columnar microstructure.
Evidence of diagenesis has been recognized in some poorly preserved areas, because the original crystallo-chemical features have been obliterated. These areas, regardless of luminescence under CL (i.e., NL or SL), show neomorphism, a random crystallographic distribution, and a disperse geochemical composition. It should be highlighted that the luminescence under fluorescence microscopy, identified in some well-preserved areas as growth lines, corresponds mainly to preserved organic-matrix remains.
Well-preserved areas of secondary and tertiary layers exhibit a well-constrained crystallographic arrangement, with the c-axis oriented parallel to the elongation axis of the prisms and perpendicular to the laminae. Three structural regions were identified in the ventral valve, based on the curvature and thickness changes in the ventral valve: the umbonal (U), thick (Tk) and thin (T) regions.
Furthermore, the geochemical composition of these areas is well-constrained in MTE (Ca, Mg, Sr, Na, S, Mn, and Fe) as well as in isotope composition (δ13C, δ18O and 87Sr/86Sr), with a characteristic variability possibly related to shell growth-rate differences, these being identified as geochemical ‘vital effects’. Younger regions with higher growth rates (e.g., umbonal and the secondary—tertiary-layer interphase) are Mg-, S- and Na-enriched and δ13C- and δ18O-depleted, whereas older zones (the inner zones of the Tk- and T-regions) are Mg-, S- and Na-impoverished and δ13C- and δ18O-enriched. These latter regions are the most suitable areas for geochemical analysis of the Gigantoproductus.
The ‘vital effects’ caused by kinetic effects have a direct influence on the paleotemperature estimation; up until ~1.05‰ of δ18O and 14 mmol/mol of Mg/Ca, intra-shell variation was detected, which can modify the ulterior palaeoclimatological interpretation, underestimating or overestimating the environmental conditions of the LPIA.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/life13030714/s1, Figure S1: Morphological characters and figure of Gigantoproductus sp.; Figure S2: CIP analysis of thin region, umbonal region, and thick region of Gigantoproductus shells. Figure S3: Boxplot of the seven analyzed elements in well-preserved and poorly-preserved areas; Figure S4: All drill locations sampled in the specimens over a brachiopod scheme to compare its relative position into the ventral valve; Table S1: Major, minor and trace elements data in ppm from Gigantoproductus sp.

Author Contributions

Conceptualization, I.C., S.R. and J.R.M.-C.; methodology, J.R.M.-C., J.A.C. and I.C.; investigation, J.R.M.-C., I.C., J.A.C., P.C., E.F.-M. and S.R.; writing—original draft preparation, I.C. and J.R.M.-C.; writing—review and editing, J.R.M.-C., I.C., J.A.C., P.C., E.F.-M. and S.R.; visualization, J.R.M.-C. and I.C.; supervision, I.C., P.C. and S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are included in figures and tables of this article.

Acknowledgments

Financial support through the Spanish Ministerio de Economía y Competitividad (research project CGL2016-78738-P) and the Complutense University Research Group (910231) is gratefully acknowledged. J.R.M.C. Acknowledges financial support through an FPI-MINECO grant. This article is a contribution to the Spanish Working Group IGCP 596 (UNESCO). We thank the anonymous reviewers and the editors for their help and constructive comments on the text.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Isbell, J.L.; Henry, L.C.; Gulbranson, E.L.; Limarino, C.O.; Fraiser, M.L.; Koch, Z.J.; Ciccioli, P.L.; Dineen, A.A. Glacial paradoxes during the late Paleozoic ice age: Evaluating the equilibrium line altitude as a control on glaciation. Gondwana Res. 2012, 22, 1–19. [Google Scholar] [CrossRef]
  2. Saltzman, M.R. Late Paleozoic Ice Age: Oceanic Gateway or pCO2? Geology 2003, 31, 151–154. [Google Scholar] [CrossRef]
  3. Powell, M.G. Climatic basis for sluggish macroevolution during the late Paleozoic ice age. Geology 2005, 33, 381–384. [Google Scholar] [CrossRef]
  4. Fielding, C.R.; Frank, T.D.; Birgenheier, L.P.; Rygel, M.C.; Jones, A.T.; Roberts, J. Stratigraphic imprint of the Late Palaeozoic Ice Age in eastern Australia: A record of alternating glacial and nonglacial climate regime. J. Geol. Soc. Lond. 2008, 165, 129–140. [Google Scholar] [CrossRef]
  5. Shi, G.R.; Waterhouse, J.B. Late Palaeozoic global changes affecting high-latitude environments and biotas: An introduction. Palaeogeogr. Palaeoclim. Palaeoecol. 2010, 298, 1–16. [Google Scholar] [CrossRef]
  6. Montañez, I.P.; Poulsen, C.J. The Late Paleozoic Ice Age: An Evolving Paradigm. Annu. Rev. Earth Planet. Sci. 2013, 41, 629–656. [Google Scholar] [CrossRef] [Green Version]
  7. Qiao, L.; Shen, S.-Z. Global paleobiogeography of brachiopods during the Mississippian—Response to the global tectonic reconfiguration, ocean circulation, and climate changes. Gondwana Res. 2014, 26, 1173–1185. [Google Scholar] [CrossRef]
  8. Goddéris, Y.; Donnadieu, Y.; Carretier, S.; Aretz, M.; Dera, G.; Macouin, M.; Regard, V. Onset and ending of the late Palaeozoic ice age triggered by tectonically paced rock weathering. Nat. Geosci. 2017, 10, 382–386. [Google Scholar] [CrossRef]
  9. Armendáriz, M.; Rosales, I.; Quesada, C. Oxygen isotope and Mg/Ca composition of Late Viséan (Mississippian) brachiopod shells from SW Iberia: Palaeoclimatic and palaeogeographic implications in northern Gondwana. Palaeogeogr. Palaeoclim. Palaeoecol. 2008, 268, 65–79. [Google Scholar] [CrossRef]
  10. McGhee, G.R.; Sheehan, P.M.; Bottjer, D.J.; Droser, M.L. Ecological ranking of Phanerozoic biodiversity crises: The Serpukhovian (early Carboniferous) crisis had a greater ecological impact than the end-Ordovician. Geology 2012, 40, 147–150. [Google Scholar] [CrossRef]
  11. Stanley, S.M.; Powell, M.G. Depressed rates of origination and extinction during the late Paleozoic ice age: A new state for the global marine ecosystem. Geology 2003, 31, 877–880. [Google Scholar] [CrossRef]
  12. Brand, U. The oxygen and carbon isotope composition of Carboniferous fossil components: Sea-water effects. Sedimentology 1982, 29, 139–147. [Google Scholar] [CrossRef]
  13. Grossman, E.L.; Yancey, T.E.; Jones, T.E.; Bruckschen, P.; Chuvashov, B.; Mazzullo, S.J.; Mii, H.-S. Glaciation, aridification, and carbon sequestration in the Permo-Carboniferous: The isotopic record from low latitudes. Palaeogeogr. Palaeoclim. Palaeoecol. 2008, 268, 222–233. [Google Scholar] [CrossRef]
  14. Stephenson, M.; Angiolini, L.; Cózar, P.; Jadoul, F.; Leng, M.; Millward, D.; Chenery, S. Northern England Serpukhovian (early Namurian) farfield responses to southern hemisphere glaciation. J. Geol. Soc. 2010, 167, 1171–1184. [Google Scholar] [CrossRef] [Green Version]
  15. Brand, U.; Jiang, G.; Azmy, K.; Bishop, J.; Montañez, I.P. Diagenetic evaluation of a Pennsylvanian carbonate succession (Bird Spring Formation, Arrow Canyon, Nevada, U.S.A.)—1: Brachiopod and whole rock comparison. Chem. Geol. 2012, 308-309, 26–39. [Google Scholar] [CrossRef]
  16. Roark, A.; Grossman, E.L.; Lebold, J. Low seasonality in central equatorial Pangea during a late Carboniferous highstand based on high-resolution isotopic records of brachiopod shells. Bull. Geol. Soc. Am. 2016, 128, 597–608. [Google Scholar] [CrossRef]
  17. Popp, B.; Anderson, T.F.; Sandberg, P.A.B.N. Brachiopods as indicators of original isotopic compositions in some Paleozoic limestones. Geol. Soc. Am. Bull. 1986, 97, 1262–1269. [Google Scholar] [CrossRef]
  18. Banner, J.L.; Kaufman, J. The isotopic record of ocean chemistry and diagenesis preserved in non-luminescent brachiopods from Mississippian carbonate rocks, Illinois and Missouri. Geol. Soc. Am. Bull. 1994, 106, 1074–1082. [Google Scholar] [CrossRef]
  19. Bruckschen, P.; Bruhn, F.; Veizer, J.; Buhl, D. 87Sr86Sr isotopic evolution of Lower Carboniferous seawater: Dinantian of western Europe. Sediment. Geol. 1995, 100, 63–81. [Google Scholar] [CrossRef]
  20. Bruckschen, P.; Oesmann, S.; Veizer, J. Isotope stratigraphy of the European Carboniferous: Proxy signals for ocean chemistry, climate and tectonics. Chem. Geol. 1999, 161, 127–163. [Google Scholar] [CrossRef]
  21. Bruckschen, P.; Veizer, J. Oxygen and carbon isotopic composition of Dinantian brachiopods: Paleoenvironmental implications for the Lower Carboniferous of western Europe. Palaeogeogr. Palaeoclim. Palaeoecol. 1997, 132, 243–264. [Google Scholar] [CrossRef]
  22. Mii, H.-S.; Grossman, E.L.; Yancey, T.E. Carboniferous isotope stratigraphies of North America: Implications for Carboniferous paleoceanography and Mississippian glaciation. Bull. Geol. Soc. Am. 1999, 111, 960–973. [Google Scholar] [CrossRef]
  23. Grossman, E.L.; Bruckschen, P.; Mii, H.-S.; Chuvashov, B.I.; Yancey, T.E.; Veizer, J. Carboniferous Paleoclimate and Global Change: Isotopic Evidence from the Russian Platform. In Carboniferous Stratigraphy and Paleogeography in Eurasia; Texas A&M University: College Station, TX, USA, 2002; pp. 61–71. [Google Scholar]
  24. Batt, L.S.; Montañez, I.P.; Isaacson, P.; Pope, M.C.; Butts, S.H.; Abplanalp, J. Multi-carbonate component reconstruction of mid-carboniferous (Chesterian) seawater δ13C. Palaeogeogr. Palaeoclim. Palaeoecol. 2007, 256, 298–318. [Google Scholar] [CrossRef]
  25. Brand, U. Biogeochemistry of Late Paleozoic North American brachiopods and secular variation of seawater composition. Biogeochemistry 1989, 7, 159–193. [Google Scholar] [CrossRef]
  26. Veizer, J.; Fritz, P.; Jones, B. Geochemistry of brachiopods: Oxygen and carbon isotopic records of Paleozoic oceans. Geochim. Cosmochim. Acta 1986, 50, 1679–1696. [Google Scholar] [CrossRef]
  27. Veizer, J.; Bruckschen, P.; Pawellek, F.; Diener, A.; Podlaha, O.G.; Carden, G.A.F.; Jasper, T.; Korte, C.; Strauss, H.; Azmy, K.; et al. Oxygen isotope evolution of Phanerozoic seawater. Palaeogeogr. Palaeoclim. Palaeoecol. 1997, 132, 159–172. [Google Scholar] [CrossRef]
  28. Mii, H.-S.; Grossman, E.L.; Yancey, T.E.; Chuvashov, B.; Egorov, A. Isotopic records of brachiopod shells from the Russian Platform—Evidence for the onset of mid-Carboniferous glaciation. Chem. Geol. 2001, 175, 133–147. [Google Scholar] [CrossRef]
  29. Pérez-Huerta, A.; Coronado, I.; Hegna, T.A. Understanding biomineralization in the fossil record. Earth-Science Rev. 2018, 179, 95–122. [Google Scholar] [CrossRef]
  30. Muir-Wood, H.; Cooper, G.A. Morphology, Classification and Life Habits of the Productoidea (Brachiopoda); Geological Society of America: Boulder, CO, USA, 1960; ISBN 9780813710815. [Google Scholar]
  31. Ferguson, J. Some aspects of the ecology and growth of the carboniferous gigantoproductids. Proc. Yorks. Geol. Soc. 1978, 42, 41–54. [Google Scholar] [CrossRef]
  32. Nolan, L.S.P.; Angiolini, L.; Jadoul, F.; Della Porta, G.; Davies, S.J.; Banks, V.J.; Stephenson, M.H.; Leng, M.J. Sedimentary context and palaeoecology of Gigantoproductus shell beds in the Mississippian Eyam Limestone Formation, Derbyshire carbonate platform, central England. Proc. Yorks. Geol. Soc. 2017, 61, 239–257. [Google Scholar] [CrossRef]
  33. Brunton, C.H.C.; Lazarev, S.S.; Grant, R.E. A Review and New Classification of the Brachiopod Order Productida. Palaeontology 1995, 38, 915–936. [Google Scholar]
  34. Angiolini, L.; Stephenson, M.; Leng, M.J.; Jadoul, F.; Millward, D.; Aldridge, A.; Andrews, J.; Chenery, S.; Williams, G. Heterogeneity, cyclicity and diagenesis in a Mississippian brachiopod shell of palaeoequatorial Britain. Terra Nova 2012, 24, 16–26. [Google Scholar] [CrossRef]
  35. Angiolini, L.; Crippa, G.; Azmy, K.; Capitani, G.; Confalonieri, G.; Della Porta, G.; Griesshaber, E.; Harper, D.A.T.; Leng, M.J.; Nolan, L.; et al. The giants of the phylum Brachiopoda: A matter of diet? Palaeontology 2019, 62, 889–917. [Google Scholar] [CrossRef]
  36. Grossman, E.L.; Mii, H.-S.; Zhang, C.; Yancey, T.E. Chemical Variation in Pennsylvanian Brachiopod Shells--Diagenetic, Taxonomic, Microstructural, and Seasonal Effects. J. Sediment. Res. 1996, 66, 1011–1022. [Google Scholar] [CrossRef]
  37. Rush, P.F.; Chafetz, H. Fabric-retentive, Non-luminescent Brachiopods as Indicators of Original 13C and 18O Composition: A Test. J. Sediment. Petrol. 1990, 60, 968–981. [Google Scholar] [CrossRef]
  38. Pérez-Huerta, A.; Cusack, M.; Janousch, M.; Finch, A.A. Influence of crystallographic orientation of biogenic calcite on in situ Mg XANES analyses. J. Synchrotron Radiat. 2008, 15, 572–575. [Google Scholar] [CrossRef] [Green Version]
  39. Urey, H.C.; Lowenstam, H.A.; Epstein, S.; McKinney, C.R. Measurement of Paleotemperatures and Temperatures and the Southeastern United States. Bull. Geol. Soc. Am. 1951, 62, 399–416. [Google Scholar] [CrossRef]
  40. Carpenter, S.J.; Lohmann, K.C. δ18O and δ13C values of modern brachiopod shells. Geochim. Cosmochim. Acta 1995, 59, 3749–3764. [Google Scholar] [CrossRef]
  41. Brand, U.; Azmy, K.; Griesshaber, E.; Bitner, M.A.; Logan, A.; Zuschin, M.; Ruggiero, E.; Colin, P.L. Carbon isotope composition in modern brachiopod calcite: A case of equilibrium with seawater? Chem. Geol. 2015, 411, 81–96. [Google Scholar] [CrossRef]
  42. Rollion-Bard, C.; Saulnier, S.; Vigier, N.; Schumacher, A.; Chaussidon, M.; Lécuyer, C. Variability in magnesium, carbon and oxygen isotope compositions of brachiopod shells: Implications for paleoceanographic studies. Chem. Geol. 2016, 423, 49–60. [Google Scholar] [CrossRef]
  43. Rollion-Bard, C.; Milner Garcia, S.; Burckel, P.; Angiolini, L.; Jurikova, H.; Tomašových, A.; Henkel, D. Assessing the biomineralization processes in the shell layers of modern brachiopods from oxygen isotopic composition and elemental ratios: Implications for their use as paleoenvironmental proxies. Chem. Geol. 2019, 524, 49–66. [Google Scholar] [CrossRef]
  44. Riechelmann, S.; Mavromatis, V.; Buhl, D.; Dietzel, M.; Eisenhauer, A.; Immenhauser, A. Impact of diagenetic alteration on brachiopod shell magnesium isotope (δ26Mg) signatures: Experimental versus field data. Chem. Geol. 2016, 440, 191–206. [Google Scholar] [CrossRef]
  45. Pérez-Huerta, A.; Fred, T.A.C.; Andrus, F.T. Vital Effects in the Context of Biomineralization. In Workshop on Biominerals and Biomineralization Processes; Fernández-Díaz, L., Astilleros, J.M., Eds.; Sociedad Española de Mienralogía: Madrid, Spain, 2010; Volume 7, pp. 35–45. [Google Scholar]
  46. Weiner, S.; Dove, P.M. An Overview of Biomineralization Processes and the Problem of the Vital Effect. Rev. Miner. Geochem. 2003, 54, 1–29. [Google Scholar] [CrossRef]
  47. De Yoreo, J.J.; Gilbert, P.U.P.A.; Sommerdijk, N.A.J.M.; Penn, R.L.; Whitelam, S.; Joester, D.; Zhang, H.; Rimer, J.D.; Navrotsky, A.; Banfield, J.F.; et al. Crystallization by particle attachment in synthetic, biogenic, and geologic environments. Science 2015, 349, aaa6760. [Google Scholar] [CrossRef] [PubMed]
  48. Coronado, I.; Fine, M.; Bosellini, F.R.; Stolarski, J. Impact of ocean acidification on crystallographic vital effect of the coral skeleton. Nat. Commun. 2019, 10, 2896. [Google Scholar] [CrossRef] [Green Version]
  49. Auclair, A.-C.; Joachimski, M.M.; Lécuyer, C. Deciphering kinetic, metabolic and environmental controls on stable isotope fractionations between seawater and the shell of Terebratalia transversa (Brachiopoda). Chem. Geol. 2003, 202, 59–78. [Google Scholar] [CrossRef]
  50. Parkinson, D.; Curry, G.B.; Cusack, M.; Fallick, A.E. Shell structure, patterns and trends of oxygen and carbon stable isotopes in modern brachiopod shells. Chem. Geol. 2005, 219, 193–235. [Google Scholar] [CrossRef]
  51. Romanin, M.; Crippa, G.; Ye, F.; Brand, U.; Bitner, M.; Gaspard, D.; Häussermann, V.; Laudien, J. A Sampling Strategy for Recent and Fossil Brachiopods: Selecting the Optimal Shell Segment for Geochemical Analyses. Riv. Ital. Paleontol. Stratigr. 2018, 124, 343–359. [Google Scholar] [CrossRef]
  52. Bajnai, D.; Fiebig, J.; Tomašových, A.; Milner Garcia, S.; Rollion-Bard, C.; Raddatz, J.; Löffler, N.; Primo-Ramos, C.; Brand, U. Assessing kinetic fractionation in brachiopod calcite using clumped isotopes. Sci. Rep. 2018, 8, 533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Brand, U.; Logan, A.; Hiller, N.; Richardson, J. Geochemistry of modern brachiopods: Applications and implications for oceanography and paleoceanography. Chem. Geol. 2003, 198, 305–334. [Google Scholar] [CrossRef]
  54. Cusack, M.; Pérez-Huerta, A.; Janousch, M.; Finch, A.A. Magnesium in the lattice of calcite-shelled brachiopods. Chem. Geol. 2008, 257, 59–64. [Google Scholar] [CrossRef]
  55. Pérez-Huerta, A.; Cusack, M.; Jeffries, T.E.; Williams, C.T. High resolution distribution of magnesium and strontium and the evaluation of Mg/Ca thermometry in Recent brachiopod shells. Chem. Geol. 2008, 247, 229–241. [Google Scholar] [CrossRef]
  56. Buening, N.; Carlson, S.J. Geochemical investigation of growth in selected Recent articulate brachiopods. Lethaia 1992, 25, 331–345. [Google Scholar] [CrossRef]
  57. Brand, U.; Azmy, K.; Bitner, M.A.; Logan, A.; Zuschin, M.; Came, R.; Ruggiero, E. Oxygen isotopes and MgCO3 in brachiopod calcite and a new paleotemperature equation. Chem. Geol. 2013, 359, 23–31. [Google Scholar] [CrossRef]
  58. Brand, U.; Bitner, M.A.; Logan, A.; Azmy, K.; Crippa, G.; Angiolini, L.; Colin, P.; Griesshaber, E.; Harper, E.; Haussermann, V. Brachiopod-Based Oxygen-Isotope Thermometer: Update and Review. Riv. Ital. Paleontol. Stratigr. 2019, 125, 775–778. [Google Scholar]
  59. Cózar, P.; Rodríguez, S. Propuesta de nueva nomenclatura para las unidades del Carbonífero Inferior del sector norte del área del Guadiato (Córdoba). Bol. Geol. Min. 1999, 110, 237–254. [Google Scholar]
  60. Cózar, P.; Rodríguez, S. Pendleian(early Serpukhovian) marine carbonates from SW Spain: Sedimentology, biostratigraphy and depositional model. Geol. J. 2004, 39, 25–47. [Google Scholar] [CrossRef]
  61. Medina-Varea, P.; Sarmiento, G.; Rodríguez, S.; Cózar, P. Early Serpukhovian Conodonts from the Guadiato Area (Córdoba, Spain). Coloq. Paleontol. 2005, 55, 21–50. [Google Scholar]
  62. Cózar, P.; Rodríguez, S.; Somerville, I.D. Large multi-biotic cyanoliths from relatively deep-water facies in the early Serpukhovian of SW Spain. Facies 2003, 49, 31–48. [Google Scholar] [CrossRef]
  63. Cózar, P. Bioestratigrafía Con Foraminíferos Del Carbonífero Inferior Del Sector Norte Del Área Del Guadiato (Córdoba). Ph.D. Thesis, Universidad Complutense, Madrid, Spain, 1998. [Google Scholar]
  64. Cózar, P.; Rodríguez, S. Earlandiacea (Foraminiferida) del Carbonífero inferior del Área del Guadiato (suroeste de España). Rev. Soc. Española Paleontol. 2000, 15, 89–99. [Google Scholar] [CrossRef]
  65. Colmenero, J.R.; Fernandez, L.P.; Moreno, C.; Bahamonde, J.R.; Barba, P.; Heredia, N.; Gonzalez, F. Carboniferous the Geology of Spain; Gibbons, W., Moreno, T., Colmenero, J.R., Eds.; Geological Society, Bath, United Kingdom (GBR): London, UK, 2002; pp. 93–116. ISBN 1862391106. [Google Scholar]
  66. Coronado, I.; Pérez-Huerta, A.; Rodríguez, S. Primary biogenic skeletal structures in Multithecopora (Tabulata, Pennsylvanian). Palaeogeogr. Palaeoclim. Palaeoecol. 2013, 386, 286–299. [Google Scholar] [CrossRef]
  67. Heilbronner, R.; Barret, S. Image Analysis in Earth Sciences—Microstructures and Textures of Earth Materials; Springer: London, UK, 2014. [Google Scholar]
  68. Coronado, I.; Pérez-Huerta, A.; Rodríguez, S. Computer-integrated polarisation (CIP) in the analysis of fossils: A case of study in a Palaeozoic coral (Sinopora, Syringoporicae, Carboniferous). Hist. Biol. 2015, 27, 1098–1112. [Google Scholar] [CrossRef]
  69. Coronado, I.; Fernández-Martínez, E.; Rodríguez, S.; Tourneur, F. Reconstructing a Carboniferous inferred coral-alcyonarian association using a biomineralogical approach. Geobiology 2015, 13, 340–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Coronado, I.; Rodríguez, S. Biomineral structure and crystallographic arrangement of cerioid and phaceloid growth in corals belonging to the Syringoporicae (Tabulata, Devonian–Carboniferous): A genetic reflection. Geol. Mag. 2016, 153, 718–742. [Google Scholar] [CrossRef] [Green Version]
  71. Coronado, I.; Rodríguez, S. Microstructure and crystallography in Axophyllinae. Precisions on the genus Morenaphyllum. Span. J. Palaeontol. 2018, 33, 41–56. [Google Scholar] [CrossRef]
  72. Barret, S. Image Analysis and the Internet. Sci. Data Manag. 1997, 1, 18–25. [Google Scholar]
  73. Vischer, N.O.E.; Huls, P.G.; Woldringh, C.L. Object-Image—An Interactive Image-Analysis Program Using Structured Point Collection. Bin. Microbiol. 1994, 6, 160–166. [Google Scholar]
  74. Abràmofff, M.D.; Magalhães, P.J.; Ram, S.J. Image Processing with ImageJ Part II. Biophotonics Int. 2005, 11, 36–43. [Google Scholar]
  75. Pereira, M.F.; Gama, C.; dias da Silva, D.; Fuenlabrada, J.M.; Silva, J.B.; Medina, J. Isotope geochemistry evidence for Laurussian-type sources of South Portuguese Zone Carboniferous turbidites (Variscan Orogeny). Geol. Soc. Lond. Spéc. Publ. 2021, 503, 619–642. [Google Scholar] [CrossRef]
  76. Garbelli, C.; Angiolini, L.; Shen, S.-Z. Biomineralization and global change: A new perspective for understanding the end-Permian extinction. Geology 2016, 45, 19–22. [Google Scholar] [CrossRef]
  77. Garbelli, C.; Angiolini, L.; Jadoul, F.; Brand, U. Micromorphology and differential preservation of Upper Permian brachiopod low-Mg calcite. Chem. Geol. 2012, 298, 1–10. [Google Scholar] [CrossRef]
  78. Griesshaber, E.; Schmahl, W.W.; Neuser, R.; Pettke, T.; Blüm, M.; Mutterlose, J.; Brand, U. Crystallographic texture and microstructure of terebratulide brachiopod shell calcite: An optimized materials design with hierarchical architecture. Am. Miner. 2007, 92, 722–734. [Google Scholar] [CrossRef]
  79. Casella, L.A.; He, S.; Griesshaber, E.; Fernández-Díaz, L.; Greiner, M.; Harper, E.M.; Jackson, D.J.; Ziegler, A.; Mavromatis, V.; Dietzel, M.; et al. Hydrothermal alteration of aragonitic biocarbonates: Assessment of micro- and nanostructural dissolution–reprecipitation and constraints of diagenetic overprint from quantitative statistical grain-area analysis. Biogeosciences 2018, 15, 7451–7484. [Google Scholar] [CrossRef] [Green Version]
  80. Cusack, M.; Parkinson, D.; Pérez-Huerta, A.; England, J.; Curry, G.B.; Fallick, A.E. Relationship between δ18O and minor element composition of Terebratalia transversa. Earth Environ. Sci. Trans. R. Soc. Edinb. 2007, 98, 443–449. [Google Scholar] [CrossRef]
  81. Pérez-Huerta, A.; Reed, H. Preliminary assessment of coupling the analysis of shell microstructures and microtextures as palaeoecological indicator in fossil brachiopods. Span. J. Palaeontol. 2018, 33, 129–138. [Google Scholar] [CrossRef] [Green Version]
  82. Goetz, A.J.; Griesshaber, E.; Neuser, R.D.L.; Harper, E.; Schmahl, W.W. Calcite morphology, texture and hardness in the distinct layers of rhynchonelliform brachiopod shells. Eur. J. Miner. 2009, 21, 303–315. [Google Scholar] [CrossRef]
  83. Schmahl, W.W. The Hierarchical Organization in Biomaterials: From Nanoparticles via Mesocrystals to Functionality Citation. In Workshop on Biominerals and Biomineralization Processes; Fernández Díaz, L., Astilleros García-Monge, J.M., Eds.; Sociedad Española de Mineralogía: Madrid, Spain, 2010; Volume 7, pp. 5–21. [Google Scholar]
  84. Gilbert, P.U.P.A.; Porter, S.M.; Sun, C.-Y.; Xiao, S.; Gibson, B.M.; Shenkar, N.; Knoll, A.H. Biomineralization by particle attachment in early animals. Proc. Natl. Acad. Sci. USA 2019, 116, 17659–17665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Cusack, M.; Dauphin, Y.; Chung, P.; Pérez-Huerta, A.; Cuif, J.-P. Multiscale structure of calcite fibres of the shell of the brachiopod Terebratulina retusa. J. Struct. Biol. 2008, 164, 96–100. [Google Scholar] [CrossRef]
  86. Cusack, M.; Pérez-Huerta, A.; Chung, P.; Parkinson, D.; Dauphin, Y.; Cuif, J.-P. Oxygen isotope equilibrium in brachiopod shell fibres in the context of biological control. Miner. Mag. 2008, 72, 239–242. [Google Scholar] [CrossRef] [Green Version]
  87. Grossman, E.; Mii, H.-S.; Yancey, T.E. Stable isotopes in Late Pennsylvanian brachiopods from the United States: Implications for Carboniferous paleoceanography. Geol. Soc. Am. Bull. 1993, 105, 1284–1296. [Google Scholar] [CrossRef]
  88. Long, J.V.P.; Agrell, S.O. The cathodo-luminescence of minerals in thin section. Miner. Mag. J. Miner. Soc. 1965, 34, 318–326. [Google Scholar] [CrossRef]
  89. Frank, J.R.; Carpenter, A.B.; Oglesby, T.W. Cathodoluminescence and Composition of Calcite Cement in the Taum Sauk Limestone (Upper Cambrian), Southeast Missouri. J. Sediment. Res. 1982, 52, 631–638. [Google Scholar] [CrossRef]
  90. Machel, H.-G. Cathodoluminescence in Calcite and Dolomite and Its Chemical Interpretation. Geosci. Can. 1985, 12, 139–147. [Google Scholar]
  91. Boggs, S.; Krinsley, D. Application of Cathodoluminescence Imaging to the Study of Sedimentary Rocks; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  92. Barbin, V.; Gaspard, D. Cathodoluminescence of recentarticulate brachiopod shells. Implications for growth stages and diagenesis evaluation. Geobios 1995, 28, 39–45. [Google Scholar] [CrossRef]
  93. Barbin, V. Cathodoluminescence of Carbonate Shells: Biochemical vs Diagenetic Process. In Cathodoluminescence in Geosciences; Pagel, M., Barbin, V., Blanc, P., Ohnenstetter, D., Eds.; Springer: Berlin/Heidelberg, Germany, 2000; pp. 303–329. ISBN 978-3-642-08526-0. [Google Scholar]
  94. Barbin, V. Application of cathodoluminescence microscopy to recent and past biological materials: A decade of progress. Miner. Pet. 2013, 107, 353–362. [Google Scholar] [CrossRef]
  95. Tomašových, A.; Farkaš, J. Cathodoluminescence of Late Triassic terebratulid brachiopods: Implications for growth patterns. Palaeogeogr. Palaeoclim. Palaeoecol. 2005, 216, 215–233. [Google Scholar] [CrossRef]
  96. Benito, M.I.; Reolid, M. Belemnite taphonomy (Upper Jurassic, Western Tethys) part II: Fossil–diagenetic analysis including combined petrographic and geochemical techniques. Palaeogeogr. Palaeoclim. Palaeoecol. 2012, 358-360, 89–108. [Google Scholar] [CrossRef]
  97. Lutz, R.A.; Rhoads, D.C. Anaerobiosis and a Theory of Growth Line Formation. Science 1977, 198, 1222–1227. [Google Scholar] [CrossRef] [PubMed]
  98. Hughes, W.W.; Rosenberg, G.D.; Tkachuck, R.D. Growth increments in the shell of the living brachiopod Terebratalia transversa. Mar. Biol. 1988, 98, 511–518. [Google Scholar] [CrossRef]
  99. Curry, G. Shell Growth and Ecology of the Recent Brachiopod Terebratulina from Scotland. Palaeontology 1982, 25, 227–246. [Google Scholar]
  100. Müller, T.; Tomašových, A.; Correa, M.L.; Mertz-Kraus, R.; Mikuš, T. Mapping intrashell variation in Mg/Ca of brachiopods to external growth lines: Mg enrichment corresponds to seasonal growth slowdown. Chem. Geol. 2022, 593, 120758. [Google Scholar] [CrossRef]
  101. Busenberg, E.; Niel Plummer, L. Kinetic and thermodynamic factors controlling the distribution of SO32− and Na+ in calcites and selected aragonites. Geochim. Cosmochim. Acta 1985, 49, 713–725. [Google Scholar] [CrossRef]
  102. Rollion-Bard, C.; Blamart, D. Possible controls on Li, Na, and Mg incorporation into aragonite coral skeletons. Chem. Geol. 2015, 396, 98–111. [Google Scholar] [CrossRef]
  103. Ishikawa, M.; Ichikuni, M. Uptake of sodium and potassium by calcite. Chem. Geol. 1984, 42, 137–146. [Google Scholar] [CrossRef]
  104. Yoshimura, T.; Suzuki, A.; Tamenori, Y.; Kawahata, H. Micro-X-ray fluorescence-based comparison of skeletal structure and P, Mg, Sr, O and Fe in a fossil of the cold-water coral Desmophyllum sp., NW Pacific. Geo-Marine Lett. 2014, 34, 1–9. [Google Scholar] [CrossRef]
  105. Yoshimura, T.; Tamenori, Y.; Suzuki, A.; Kawahata, H.; Iwasaki, N.; Hasegawa, H.; Nguyen, L.T.; Kuroyanagi, A.; Yamazaki, T.; Kuroda, J.; et al. Altervalent substitution of sodium for calcium in biogenic calcite and aragonite. Geochim. Cosmochim. Acta 2017, 202, 21–38. [Google Scholar] [CrossRef]
  106. Mii, H.-S.; Grossman, E.L. Late Pennsylvanian seasonality reflected in the 18O and elemental composition of a brachiopod shell. Geology 1994, 22, 661. [Google Scholar] [CrossRef]
  107. Mucci, A. The solubility of calcite and aragonite in seawater at various salinities, temperatures, and one atmosphere total pressure. Am. J. Sci. 1983, 283, 780–799. [Google Scholar] [CrossRef]
  108. McConnaughey, T. 13C and 18O isotopic disequilibrium in biological carbonates: I. Patterns. Geochim. Cosmochim. Acta 1989, 53, 151–162. [Google Scholar] [CrossRef]
Figure 1. (A) Synthetic map showing the distribution of the main Carboniferous outcrops of the Iberian Peninsula, modified from Colmenero et al. [65]. (B) Stratigraphic log of the San Antonio section with geographic coordinates 5°8′20″ W–38°14′ N, modified from Cózar [63] and Cózar et al. [62]. (C) Early Carboniferous outcrops (upper Tournaisian to Arnsbergian) in Sierra Morena (upper part) and synthetic map with the main geological units of the Guadiato Area (lower part). Modified from Cózar [63]). Cantabrian Zone, C.Z.; West-Asturian-Leonese Zone, W.A.L.Z.; Galicia-Tras os Montes, G.T.M.Z.; Central-Iberian Zone, C.I.Z.; Ossa-Morena Zone, O. M. Z. and South Portuguese Zone, S.P.Z.).
Figure 1. (A) Synthetic map showing the distribution of the main Carboniferous outcrops of the Iberian Peninsula, modified from Colmenero et al. [65]. (B) Stratigraphic log of the San Antonio section with geographic coordinates 5°8′20″ W–38°14′ N, modified from Cózar [63] and Cózar et al. [62]. (C) Early Carboniferous outcrops (upper Tournaisian to Arnsbergian) in Sierra Morena (upper part) and synthetic map with the main geological units of the Guadiato Area (lower part). Modified from Cózar [63]). Cantabrian Zone, C.Z.; West-Asturian-Leonese Zone, W.A.L.Z.; Galicia-Tras os Montes, G.T.M.Z.; Central-Iberian Zone, C.I.Z.; Ossa-Morena Zone, O. M. Z. and South Portuguese Zone, S.P.Z.).
Life 13 00714 g001
Figure 2. (ae) Microstructural features of Gigantoproductus sp. Shells under petrographic microscopy, cross-polarized micrographs; (a,b,df), and cross-polarized with lambda plate micrographs. (c) Scheme of the shell with underscored shell regions. On the right, a line which divides the secondary and tertiary layer into the thick region (micrographs (b,d,f)). (a) The laminar microstructure of the dorsal valve with micritized patches (upper part), micrite infilling between the dorsal and the ventral valves, and inner region of the ventral valve (lower part). Scale bar 2 mm. (b) Innermost part of the thick region, showing ventral valve with columnar microstructure and several growth lines. Scale bar 1 mm. (b1) Inset showing a laminar growth line. Scale bar 100 µm. (b2) Inset showing a diffuse growth line. Scale bar 100 µm. (b3) Inset showing pseudopunctae in a growth line. Scale bar 100 µm. (d) Umbonal region of the ventral valve with several growth lines and a chaotic crystal arrangement. Scale bar 1 mm. (e) The thick region of the ventral valve with the largest columnar crystals and several growth lines. Scale bar 1 mm. (f) Thin region of the ventral valve with columnar microstructure and several growth lines. Scale bar 1 mm.
Figure 2. (ae) Microstructural features of Gigantoproductus sp. Shells under petrographic microscopy, cross-polarized micrographs; (a,b,df), and cross-polarized with lambda plate micrographs. (c) Scheme of the shell with underscored shell regions. On the right, a line which divides the secondary and tertiary layer into the thick region (micrographs (b,d,f)). (a) The laminar microstructure of the dorsal valve with micritized patches (upper part), micrite infilling between the dorsal and the ventral valves, and inner region of the ventral valve (lower part). Scale bar 2 mm. (b) Innermost part of the thick region, showing ventral valve with columnar microstructure and several growth lines. Scale bar 1 mm. (b1) Inset showing a laminar growth line. Scale bar 100 µm. (b2) Inset showing a diffuse growth line. Scale bar 100 µm. (b3) Inset showing pseudopunctae in a growth line. Scale bar 100 µm. (d) Umbonal region of the ventral valve with several growth lines and a chaotic crystal arrangement. Scale bar 1 mm. (e) The thick region of the ventral valve with the largest columnar crystals and several growth lines. Scale bar 1 mm. (f) Thin region of the ventral valve with columnar microstructure and several growth lines. Scale bar 1 mm.
Life 13 00714 g002
Figure 3. SEM images of polished shell surfaces (ad). SEM images of natural breakage of the shell (eh). (a) Innermost part of the tertiary layer. Scale bar 200 μm. (b) Middle part of the shell. Scale bar 200 μm. (c) Outermost part of the tertiary layer and secondary layer. Scale bar 200 μm. (d) Submicrometric laminae preserved in a columnar crystal. Scale bar 10 µm. (e) Laminar microstructure composed of lath crystals. Scale bar 10 µm. (f) Nanogranular texture in some columnar crystals. Scale bar 2 µm. (g) Columnar crystals of the tertiary layer; notice the massive appearance. Scale bar 100 μm. (h) Contact between columnar and laminar microstructure (dashed line) in a growth line with a granular appearance of laminae. Scale bar 10 µm. Trt: tertiary layer, Snd: secondary layer.
Figure 3. SEM images of polished shell surfaces (ad). SEM images of natural breakage of the shell (eh). (a) Innermost part of the tertiary layer. Scale bar 200 μm. (b) Middle part of the shell. Scale bar 200 μm. (c) Outermost part of the tertiary layer and secondary layer. Scale bar 200 μm. (d) Submicrometric laminae preserved in a columnar crystal. Scale bar 10 µm. (e) Laminar microstructure composed of lath crystals. Scale bar 10 µm. (f) Nanogranular texture in some columnar crystals. Scale bar 2 µm. (g) Columnar crystals of the tertiary layer; notice the massive appearance. Scale bar 100 μm. (h) Contact between columnar and laminar microstructure (dashed line) in a growth line with a granular appearance of laminae. Scale bar 10 µm. Trt: tertiary layer, Snd: secondary layer.
Life 13 00714 g003
Figure 4. CIP analysis of thin regions of Gigantoproductus ventral valves (a,d) and thick region (g). Orientation images (b,e,h). Petrographic micrographs of the studied areas (a,d,g). Standard colour lookup table (CLUT) and pole figures (c,f,i). CLUT of the orientation images (1, 5, 9). Pole figures (2–4, 6–8, 10–12). (2) Corresponds to the complete area (b), showing a pole maximum with c-axis which is strictly oriented perpendicular to the growth direction of Gigantoproductus shell. (3–4) Corresponds to selected areas. (3) Columnar area with random c-axis orientations due to recrystallisation. (4) Equivalent area to (3), showing a well-constrained c-axis orientation. (6) Corresponds to the complete area (d). (7–8) Corresponds to selected areas. (7) Fracture with degrading neomorphism, which preserves the original crystallographic orientation. (8) Equivalent area to (7), showing a better constrained c-axis orientation. (10) Corresponds to the complete area (g), showing a pole maximum, with c-axis which is strictly oriented perpendicular to the growth direction of the Gigantoproductus shell. (11–12) Corresponds to selected areas. (11) Prismatic area with random c-axis orientations, due to recrystallisation. (12) Equivalent area to (11), showing a well-constrained c-axis orientation. Pole figures were calculated as an orientation-distribution function and are provided in multiples of uniform distribution intervals of 0.5 for c-axis orientations of the complete area. Red points correspond to the punctual c-axis orientation of each pixel of selected areas.
Figure 4. CIP analysis of thin regions of Gigantoproductus ventral valves (a,d) and thick region (g). Orientation images (b,e,h). Petrographic micrographs of the studied areas (a,d,g). Standard colour lookup table (CLUT) and pole figures (c,f,i). CLUT of the orientation images (1, 5, 9). Pole figures (2–4, 6–8, 10–12). (2) Corresponds to the complete area (b), showing a pole maximum with c-axis which is strictly oriented perpendicular to the growth direction of Gigantoproductus shell. (3–4) Corresponds to selected areas. (3) Columnar area with random c-axis orientations due to recrystallisation. (4) Equivalent area to (3), showing a well-constrained c-axis orientation. (6) Corresponds to the complete area (d). (7–8) Corresponds to selected areas. (7) Fracture with degrading neomorphism, which preserves the original crystallographic orientation. (8) Equivalent area to (7), showing a better constrained c-axis orientation. (10) Corresponds to the complete area (g), showing a pole maximum, with c-axis which is strictly oriented perpendicular to the growth direction of the Gigantoproductus shell. (11–12) Corresponds to selected areas. (11) Prismatic area with random c-axis orientations, due to recrystallisation. (12) Equivalent area to (11), showing a well-constrained c-axis orientation. Pole figures were calculated as an orientation-distribution function and are provided in multiples of uniform distribution intervals of 0.5 for c-axis orientations of the complete area. Red points correspond to the punctual c-axis orientation of each pixel of selected areas.
Life 13 00714 g004
Figure 5. Fossil preservation. (ae) Cross-polarized with lambda plate micrographs (left) and the same area under cathodoluminescence (right); (fj) cathodoluminescence micrographs in other shell regions. (a) Non-luminescent well-preserved area under CL. (b) Well-preserved area with luminescence concentrated in growth lines. (c) Non-luminescent poorly preserved area with a recrystallisation patch in the center, evidence of recrystallisation. NL fracture (left arrow) (d) SL-luminescent poorly preserved area, with recrystallisation by fractures. (e) Highly luminescent poorly preserved area with fractures. (f) CL image showing a laminar growth line with the equivalent luminescent areas as in the fluorescent image. (g) Alteration patch in the ventral valve with altered growth lines, more luminescent than well-preserved areas. Luminescent sparite in the right of the image. (h) The altered external edge of the shell, more luminescent than the central and inner parts of the ventral valve. (i) Degrading neomorphism in the shell edge, with decreasing luminescence towards the interior. (j) Encrusting sponge over Gigantoproductus shell: notice the luminescent burrow (arrow). NL: non-luminescent, SL: slightly-luminescent, L: luminescent S: sparite.
Figure 5. Fossil preservation. (ae) Cross-polarized with lambda plate micrographs (left) and the same area under cathodoluminescence (right); (fj) cathodoluminescence micrographs in other shell regions. (a) Non-luminescent well-preserved area under CL. (b) Well-preserved area with luminescence concentrated in growth lines. (c) Non-luminescent poorly preserved area with a recrystallisation patch in the center, evidence of recrystallisation. NL fracture (left arrow) (d) SL-luminescent poorly preserved area, with recrystallisation by fractures. (e) Highly luminescent poorly preserved area with fractures. (f) CL image showing a laminar growth line with the equivalent luminescent areas as in the fluorescent image. (g) Alteration patch in the ventral valve with altered growth lines, more luminescent than well-preserved areas. Luminescent sparite in the right of the image. (h) The altered external edge of the shell, more luminescent than the central and inner parts of the ventral valve. (i) Degrading neomorphism in the shell edge, with decreasing luminescence towards the interior. (j) Encrusting sponge over Gigantoproductus shell: notice the luminescent burrow (arrow). NL: non-luminescent, SL: slightly-luminescent, L: luminescent S: sparite.
Life 13 00714 g005
Figure 6. Box plot (ad) showing the variation in Ca, Mg, Na and S (in ppm) for the proposed shell regions in Gigantoproductus ventral valve. U: umbonal region, Tk: thick region. T: thin region. Ca: calcium, Mg: magnesium, Na: sodium, S: sulphur. Outliers have been removed.
Figure 6. Box plot (ad) showing the variation in Ca, Mg, Na and S (in ppm) for the proposed shell regions in Gigantoproductus ventral valve. U: umbonal region, Tk: thick region. T: thin region. Ca: calcium, Mg: magnesium, Na: sodium, S: sulphur. Outliers have been removed.
Life 13 00714 g006
Figure 7. Mg/Ca, S/Ca, Na/Ca transects from the shell edge to the shell interior in well-preserved areas (a,b) and poorly preserved area (c).
Figure 7. Mg/Ca, S/Ca, Na/Ca transects from the shell edge to the shell interior in well-preserved areas (a,b) and poorly preserved area (c).
Life 13 00714 g007
Figure 8. (a,b) δ13C vs. δ18O diagram of 11 Gigantoproductus shells. Well-preserved areas (filled-in black), dorsal valve (blue circles) and poorly preserved areas (red diamonds). (b) Values of well-preserved areas of the tertiary layer in each region. Plan of the shell with underscored shell regions. U: umbonal region; Tk: thick region; T: thin region; DV: Dorsal Valve.
Figure 8. (a,b) δ13C vs. δ18O diagram of 11 Gigantoproductus shells. Well-preserved areas (filled-in black), dorsal valve (blue circles) and poorly preserved areas (red diamonds). (b) Values of well-preserved areas of the tertiary layer in each region. Plan of the shell with underscored shell regions. U: umbonal region; Tk: thick region; T: thin region; DV: Dorsal Valve.
Life 13 00714 g008
Table 1. Descriptive statistics of major, minor and trace-element (MTE) composition of five preservation states identified. Values in ppm.
Table 1. Descriptive statistics of major, minor and trace-element (MTE) composition of five preservation states identified. Values in ppm.
Preservation StateElementN totalMeanS.D.MinimumMedianMaximum
WP-NLCa205390,449.96880.8372,632.4391,045.6406,978.5
Mg3380.5911.31218.33238.65886.3
Sr1737.9430.5642.71767.32968.1
Mn155.3214.30.00.0836.6
Fe184.3263.00.00.01026.2
S1060.2328.8388.5993.21894.4
Na1038.2358.40.01031.02202.8
PP-NLCa20372,422.629,616.5269,479.6380,938.4400,052.1
Mg4651.71876.9645.35020.87273.4
Sr1763.1473.2465.11839.22443.8
Mn278.9467.20.027.11828.1
Fe955.41710.00.0427.66499.1
S1490.5929.90.01620.03576.5
Na926.0574.80.0997.62017.4
WP-SLCa44382,121.08596.2365,870.4382,139.2398,808.4
Mg4164.81044.62291.83959.46748.7
Sr2208.9444.71166.92262.03027.2
Mn314.9367.30.0170.41332.3
Fe371.2408.20.0268.21391.5
S994.7373.6460.6979.21974.5
Na932.9358.0311.5867.81713.3
PP-SLCa80378,873.3141,25.2313,975.9383,615.3398,608.2
Mg4378.51468.81724.94010.69661.7
Sr1975.1501.0879.41894.13729.1
Mn125.0220.30.00.01038.0
Fe451.7880.70.0139.96499.1
S1202.5565.04.01125.43576.5
Na920.4378.329.7934.52010.0
PP-LCa15382,641.08793.2358,236.3383,947.7394,598.2
Mg4829.21635.32249.64589.67786.0
Sr1981.5473.81217.72232.42765.1
Mn582.0292.4302.1457.01332.3
Fe56.089.60.00.0287.6
S1264.5462.3596.71093.41918.4
Na966.7276.5504.41023.51550.2
WP-NL: well-preserved, non-luminescent; PP-NL: poorly preserved non-luminescent; WP-SL: well-preserved, slightly luminescent; PP-SL: poorly preserved, slightly luminescent; PP-L: poorly preserved, luminescent; S.D. Standard deviation.
Table 2. Descriptive statistics of major, minor and trace-element (MTE) composition of tertiary-layer shell regions and secondary layer. Values in ppm.
Table 2. Descriptive statistics of major, minor and trace-element (MTE) composition of tertiary-layer shell regions and secondary layer. Values in ppm.
Shell PartElementN TotalMeanS.D.MinimumMedianMaximum
UCa33388,430.16175.9373,604.5389,194.3398,379.5
Mg3259.7840.41899.83178.35735.5
Sr1604.6551.8642.71648.92570.6
Mn227.5255.80170.4836.6
Fe205266.8038.91026.2
S930.1303.3440.6893.11786.2
Na1131.9337.6526.61105.11995.2
TkCa97393,264.26774.8372,632.4394,433.8406,978.5
Mg3012.1722.71218.32943.14999.7
Sr1656.4408.8845.61623.62460.7
Mn140.9202.100774.6
Fe219.4282.1046.6925.1
S1040.1371.3408.5957.22367
Na1063.5401.9010682202.8
TCa90386,561.57115.3366,270.7387,503.8397,164.3
Mg4126.51125.42056.64067.97273.4
Sr1890.7330.41107.71902.62968.1
Mn160.1244.8001332.3
Fe161.4264.3001041.7
S1249.2470.2224.31165.53576.5
Na968.8346.50945.71579.8
Secondary layerCa30379,749.48141.1365,870.4378,940.5398,808.4
Mg4444.01076.22780.34565.56748.7
Sr2283.4360.21462.92367.73027.2
Mn182.3210.10.0116.2704.9
Fe370.0453.70.0225.41609.2
S980.8288.5460.61039.31549.9
Na1013.9385.7311.5997.51713.3
U: umbonal region; Tk: thick region; T: thin region; S.D. Standard deviation.
Table 3. Isotopic composition of the structural regions identified in the tertiary layer of Gigantoproductus shells and poorly preserved areas. Bold highlights the mean of each structural part.
Table 3. Isotopic composition of the structural regions identified in the tertiary layer of Gigantoproductus shells and poorly preserved areas. Bold highlights the mean of each structural part.
Shell Regionδ18Oδ13C87Sr/86Sr
(Std Err)
M.S.D.Shell Regionδ18Oδ13C87Sr/86Sr
(Std Err)
M.S.D.Shell Regionδ18Oδ13C87Sr/86Sr
(Std Err)
M.S.D.
Umbornal−3.42.78 Thin−2.913.350.70786
−3.612.35 −2.773.41(4−6) Poorly preserved−3.711.84
−3.782.39 −3.133.290.707845 −2.864.27
−3.552.68 −2.992.95(4−6) −5.063.620.707893
−3.252.52 −2.933.15 −4.123.68(4−6)
−3.142.97 −2.333.04 −2.323.87
−3.272.81 −2.443.17 −6.283.08
Mean−3.432.64 0.22/0.23−2.343.38 −4.371.18
Thick−2.822.180.70783 −2.543.29 −1.643.05
−3.392.1(4−6) −2.272.89 −1.893.76
−2.622.25 Mean−2.673.19 0.32/0.18−3.873.30.707807
−2.922.29 Dorsal Valve−3.252.49 −2.162.61(5−6)
−23.09 −4.021.54 −2.142.43
−2.583.04 −1.463.56 −1.833.46
−2.332.960.707831 −3.132.13 −4.033.34
−2.322.94(4−6) Mean−2.972.43 1.08/0.85Mean−3.313.11 1.4/0.84
−2.712.86
Mean−2.632.63 0.40/0.42
M.S.D. Mean standard deviation; Std Err: Standard deviation error.
Table 4. Summary of structural and geochemical features of the five states of preservation identified.
Table 4. Summary of structural and geochemical features of the five states of preservation identified.
FeaturesPreservation States
Well-PreservedPoorly Preserved
Tertiary LayerSecondary LayerTertiary LayerGrowth LinesDorsal ValveSecondary Layer
StructuralMicrostructurePrismatic crystals with noticeable microlaminae, preferential orientationGrowth lines: original organic-matrix remains and preferential orientation of the crystalsLath crystals, original organic-matrix remains and preferential orientationRecrystallisation (e.g., degrading neomorphism), sparite cements, filled fractures, small patches of crystals without a preferential orientationRecrystallisation, delaminationRecrystallisation, matrix infilling, micritzation, sparite cements, filled fracturesRecrystallisation, delamination, micritized patches, matrix infilling, filled fractures
CathodoluminescenceNon-luminescentSlightly luminescent, luminescence pattern in growth linesSlightly luminescent, luminescence bandingNon-luminescentIncreasing luminescence near fractured areasLuminescentSlightly luminescent
GeochemicalCrystallographicHighly arranged crystals: c-axis oriented perpendicular to shell growthHighly arranged crystals: c-axis oriented subparallel to shell growthHighly arranged crystals: c-axis oriented parallel to shell growthPoorly arranged crystals: c-axis randomly oriented
MTE (ppm)Mg: 3380, Sr: 1737, Mn: 155, Fe: 184Mg: 4164, Sr: 2208, Mn: 314, Fe: 371Mg: 4444 Sr: 2283 Mn: 182 Fe: 370Mg: 4651, Sr: 1763, Mn: 278, Fe: 955Mg: 3765, Sr: 1922, Mn: 971, Fe: 858_Mg: 4849, Sr: 2323, Mn: 117, Fe: 838
δ13C–δ18OLow dispersion of the ratio δ18O: −3.78‰ to −2‰; δ13C: 2.09‰ to 3.4‰; grouped in clusters depending on ontogeny development High dispersion of the ratio δ18O: −6.28‰ to −1.83‰; δ13C: 1.18‰ to 4.27‰
87Sr/86SrLow dispersion of the ratio: 0.707830–0.707860 High dispersion of the ratio: 0.707807–0.707893
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mateos-Carralafuente, J.R.; Coronado, I.; Cruz, J.A.; Cózar, P.; Fernández-Martínez, E.; Rodríguez, S. Diagenetic and Biological Overprints in Geochemical Signatures of the Gigantoproductus Tertiary Layer (Brachiopoda): Assessing the Paleoclimatic Interpretation. Life 2023, 13, 714. https://doi.org/10.3390/life13030714

AMA Style

Mateos-Carralafuente JR, Coronado I, Cruz JA, Cózar P, Fernández-Martínez E, Rodríguez S. Diagenetic and Biological Overprints in Geochemical Signatures of the Gigantoproductus Tertiary Layer (Brachiopoda): Assessing the Paleoclimatic Interpretation. Life. 2023; 13(3):714. https://doi.org/10.3390/life13030714

Chicago/Turabian Style

Mateos-Carralafuente, José R., Ismael Coronado, Juncal A. Cruz, Pedro Cózar, Esperanza Fernández-Martínez, and Sergio Rodríguez. 2023. "Diagenetic and Biological Overprints in Geochemical Signatures of the Gigantoproductus Tertiary Layer (Brachiopoda): Assessing the Paleoclimatic Interpretation" Life 13, no. 3: 714. https://doi.org/10.3390/life13030714

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop