Next Article in Journal
Characterization of Fine-Scale Turbulence Generated in a Laboratory Orbital Shaker and Its Influence on Skeletonema costatum
Previous Article in Journal
Binocular-Vision-Based Obstacle Avoidance Design and Experiments Verification for Underwater Quadrocopter Vehicle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Material Source of Sediments from West Clarion–Clipperton Zone (Pacific): Evidence from Rare Earth Element Geochemistry and Clay Minerals Compositions

1
Ocean College, Zhejiang University, Zhoushan 316000, China
2
Key Laboratory of Submarine Geosciences, Second Institute of Oceanography, Ministry of Natural Resources, Hangzhou 310012, China
3
Institute of Science and Environment, University of Saint Joseph, Rua de Londres 106, Macao, China
4
Department of Earth Sciences, School of Physical Sciences, University of Adelaide, Adelaide 5005, Australia
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2022, 10(8), 1052; https://doi.org/10.3390/jmse10081052
Submission received: 19 June 2022 / Revised: 16 July 2022 / Accepted: 16 July 2022 / Published: 31 July 2022
(This article belongs to the Section Geological Oceanography)

Abstract

:
The geochemistry and mineralogy of sediments provide relevant information for the understanding of the origin and metallogenic mechanism of ferromanganese nodules and crusts. At present, there are still few studies on the sediment origin of the Clarion–Clipperton Zone (CCZ) of the east Pacific, particularly on the systematic origin of sediments with a longer history/length. Here, bulk sediment geochemistry and clay mineral compositions were analyzed on a 5.7 m gravity core (GC04) obtained at the CCZ, an area rich in polymetallic nodules. The results indicate that the average total content of rare earth elements (REE), including yttrium (REY), in sediments is 454.7 ppm and the REEs distribution patterns normalized by the North American Shale Composite of samples are highly consistent, with all showing negative Ce anomalies and more obvious enrichment in heavy REE (HREE) than that of light REE (LREE). Montmorillonite/illite ratio, discriminant functions and smear slide identification indicate multiple origins for the material, and are strongly influenced by contributions from marine biomass, while terrestrial materials, seamount basalts and their alteration products and authigenic source also make certain contributions. The REY characteristics of the sediments in the study area are different from those of marginal oceanic and back-arc basins, and more similar to pelagic deep-sea sediments. Based on LREE/HREE-1/δCe and LREE/HREE-Y/Ho diagrams, we conclude that samples from the study area had pelagic sedimentary properties which suffered from a strong “seawater effect”.

1. Introduction

Deep-sea sediments refer to those deposited on the seafloor with water depths > 2000 m, which comprise a mixture of multiple geochemical components of different origins, such as terrigenous detrital materials, biological residues (biological phosphate and calcareous and siliceous biological fragments), hydrothermal (e.g., Fe–Mn-oxyhydroxides and sulfide minerals), volcanogenic materials and hydrogenetic precipitates [1,2,3,4]. Marine ferromanganese (Fe–Mn) crusts and nodules are generally formed by precipitation of Fe oxides and Mn oxides from seawater, and predominantly comprise δ-MnO2 containing FeOOH·xH2O [5,6,7], with their mineralizing substances derived from seawater [8].
Deep-sea sediments are also the sinks for some elements in the ocean, promoting the formation of Fe–Mn crusts and nodules. Most Fe–Mn crusts grow on sediment-free rock surfaces of the seamounts, while Fe–Mn nodules form on the surface of sediments by accretion around a nucleus [7]. Diagenetic nodules are formed by the precipitation of metals from sediment pore waters, under oxic or suboxic conditions [9,10]. Mn, Ni, Cu, Zn, and other trace metals are transported to the nodules after they are mobilized by diagenetic processes which occur in the suboxic and anoxic layers of the sediment [11]. Previous studies have indicated that the nodules do not simply or completely dissolve upon burial under suboxic conditions but are rather subject to alteration processes, in which nodules can act as a sink for elements (Co, W), or a source for elements (Ni, Mo) to the pore water of sediments [12].
The origin of sediments is of great relevance to explanations for the formation and metallogenic mechanism of Fe–Mn nodules and crusts. For example, due to the strong influence of terrigenous materials, South China Sea crusts/nodules had probably much higher growth rates (mean 19.20 mm/Ma) than most hydrogenetic Fe–Mn deposits in other open ocean areas and contain a sufficient supply of ore-forming terrigenous detritus [13]. By contrast, the Clarion–Clipperton Fracture Zone (CCZ) in the eastern equatorial Pacific Ocean is far away from the Asian continent and the supply of terrigenous materials to the area is limited. The growth rates of nodules in CCZ can be as slow as 0.2 mm/Ma [9], which is significantly lower than marginal sea nodules. Due to variable marine tectonic environments, marine Fe–Mn deposits that have developed in distinct sea areas differ, and the sediment properties and material sources are also different.
The elemental composition of sediments has been used for the determination of sediment origin [14]. The chemical properties of rare earth elements (REE), including yttrium (REY), are very similar and are not easily transported during weathering, transportation, and deposition of source rocks, which enables the use of REEs as provenance tracers [15,16,17]. Deep-sea sediments are multi-sourced and their REE composition characteristics are a comprehensive manifestation of multi-source components [18]. REE abundance, REE distribution curve, special REE anomalies and some important REEs parameters in marine sediments can effectively reflect the source properties and formation conditions of sediment source area [19,20,21].
Although a lot of studies have been undertaken on nodules in the CCZ area, such as their abundances [9], types [22], genesis and formation processes [10,11,12], there are still few studies on their sediment origins, and even less studies have been done on the systematic origin study of sediments with a longer history/length [23,24,25,26]. In this work, by analyzing clay mineral compositions and bulk sediment geochemistry of a gravity core (GC04) collected in the western CCZ of the Pacific Ocean, we discuss the material sources and properties of sediments in this area and compared these with sediments from the cobalt-rich crust enrichment zone in the Northwest Pacific (e.g., GC1601) and with sediments in the central Pacific (e.g., Deep Sea Drilling Project (DSDP) Site 170).

2. Geological Background

The CCZ is a well-known polymetallic nodule enrichment area [27], located in the Central Eastern Pacific Ocean, north of the equatorial high productivity zone. It is limited by the Clarion Fracture Zone to the north and the Clipperton Fracture Zone to the south, with a total length of around 7240 km [9]. The studied sample GC04 core was collected from the west of CCZ (Figure 1). The basement oceanic crust in the area is mainly basalt, resulted from the expansion of the East Pacific Rise [28]. The oceanic crust expansion direction of CCZ is northwest, which is similar to the direction of the line formed by the three points (GC1601, DSDP Site 170 and GC04) in Figure 1. The Haiyang No. 4 fault, which ran through the whole area in a near NS trending, divided the CCZ into two tectonic units. The west part is mainly dominated by volcanic terrain with old stratigraphic age up to 90 Ma [29]. Fault-block structures with most low-lying hills dominate the eastern part and its oceanic crust is relatively young, formed in the Eocene, and with a sedimentation rate of overlying sediments of a few meters per million years [30]. About half of the area was covered by flat sediments where polymetallic nodules can be found with an abundance of up to 30 kg/m2 [9].

3. Samples and Methods

3.1. Samples

The GC04 core (location, 154.74951° W, 9.50048° N; total length = 570 cm; water depth = 5144 m) was retrieved from the western CCZ of the Pacific Ocean collected by R/V No. 3 Xiangyanghong during the DY-45 cruise, 2017. The whole core had yellowish to brownish colors, showing a high clay content and biological relics (Figure 2). For the bulk geochemical analyses, the core was sampled every 5 cm, in a total of 114 samples. Compositional analyses for clay were performed in 13 samples obtained at certain intervals.

3.2. Methods

3.2.1. Smear Slide Identification

Semi-quantitative analysis of smear slide identification estimates the contents of clay, calcareous organisms, and siliceous organisms in deep-sea sediments, to determine the sediment types. A small quantity of sediment sample was put on the slide, and then a few drops of distilled water were added to the sample. After that, the sample was gently smeared and scraped with a clean toothpick, so the sample was equally distributed. After air-drying, the melted fir glue was applied evenly to the deposit. Then the sample was covered with a coverslip and any air bubbles and other substances that may interfere with naming were squeezed out as much as possible. Finally, the particle size, biological and abiotic components of each smear were identified under a polarizing microscope.

3.2.2. Chemical Analyses of Bulk Sediments

Sediment samples were freeze-dried and powdered in an agate mortar prior to major element, trace element, and loss on ignition (LOI) analyses. The powdered samples were dried at 120 °C for 8 h, and then each 0.5–1 g sample was ignited at 1000 °C for 200 min in a muffle furnace to remove the organic content. Loss on ignition (LOI) was calculated from the weight loss of the samples. Concentrations of major elements were determined using an X-ray fluorescence (XRF) spectrometer (AxiosMAX 4 kw) at the Key Laboratory of Submarine Geosciences, Ministry of Natural Resources, Hangzhou, China. The XRF analysis was conducted on glass beads made with 0.600 g of the ignited sample powder well mixed with 6.00 g of 49.75% Li2B4O7:49.75% LiBO2:0.5% LiBr flux (Canada Claisse) and fused at 1190 °C for 7 min in a Pt crucible. The typical analytical uncertainty was <5% which met the requirements of the GB/T 14506.28-2010 standard.
For REY analyses (La-Lu; Y), the 50 mg of sample powder (200 mesh) was accurately weighed and digested with HNO3 and HF, and then evaporated until dry. The residue was leached and dissolved by mixed acid (HCl:HNO3 = 4:1) and 50% HNO3 several times. The final solution was transferred to a 100 mL polyethylene bottle, added with 1 ml (Rh + Re) mixed standard solution (1 mg/L), and then concentrations were determined using Elan DRC-e ICP-MS at the Key Laboratory of Submarine Geosciences, Ministry of Natural Resources (MNR), Hangzhou, China. The typical analytical uncertainty was <10% which met the requirements of the GB/T 14506.30-2010 standard. For the above bulk sediments geochemistry analyses, BCR-2, AVG-2, and GSP-2 were used as certified reference material (CRM).

3.2.3. Clay Minerals

The clay samples were analyzed by extracting a <2 μm fraction of the sediment. Organic matter and CaCO3 were removed by 30% H2O2 and 10% HCl, respectively, and the smear method was applied to make natural air-dried clay-oriented sheets for X-ray diffraction (XRD) analysis. The qualitative analysis of clay minerals was mainly carried out via ethylene glycol vapor saturation on the same clay-oriented flakes and then tested separately by XRD at Nanjing Institute of Geology and Mineral Resources, Nanjing, China. The test conditions used for analysis were as follows: a D/max 2500 X-ray diffractometer with an operating voltage of 45 kV, an operating current of 200 mA, and a natural slice scan range of 2.5°~30° (2θ) with a scanning speed of 8°/min. The same batch of samples was tested under the same conditions. The measured diffraction data and plots were processed using the MDI Jade 6 software (California, USA).
The relevant contents of clay minerals were calculated according to the modified method from Biscaye [31]. The peak areas of three sets of characteristic peaks of four minerals—montmorillonite (17 Å), illite (10 Å), chlorite + kaolinite (7 Å)—were examined. On the glycol saturation sheet, diffraction peaks were used as the base data, and the peak areas of the corresponding characteristic peaks of each clay mineral were multiplied by their respective weighting factors. The weighting factor of montmorillonite was determined as 1, illite as 4, and chlorite + kaolinite as 2. The relative content of chlorite to kaolinite was calculated as the ratio of peak heights around 3.5 Å at about 25 °C (2θ).
The calculation formulas are as follows:
A = A1 + 4A2 + 2A3
  B1 = A1/A × 100%
 B2 = 4A2/A × 100%
     B3 = 2A3/A × V/(V + 1)100%
      B4 = 2A3/A × V/(V + 1) × 100%
where A is the total area, A1 is the area of the montmorillonite peak, A2 is the area of the illite peak, and A3 is the area of the chlorite + kaolinite peak area. B1, B2, B3 and B4 are the percentage of montmorillonite, illite, chlorite and kaolinite mineral content (%), respectively, and V is the ratio of kaolinite peak to chlorite peak height (3.5 Å).

3.2.4. Scanning Electron Microscopy

The clay mineral samples were also analyzed by field emission scanning electron microscopy (FESEM) with Energy Dispersive Spectroscopy (EDS) at Nanjing Institute of Geology and Paleontology, Chinese Academy of Sciences, Nanjing, China. FESEM images and EDS quantification were acquired using a TESCAN MAIA3 GMU with energy dispersive X-ray spectroscopy (OXFORD ULTIM MAX 170, England). The sample was uncoated and imaged in a partial vacuum using backscatter (BSE) or secondary electron detectors (SED) at an accelerating voltage of 10–20 keV. EDS analyses were conducted with the Large Area Mapping (LAM) technique using an accelerating voltage of 20 keV for an acquisition rate of around 130,000 cps/s.

4. Results

4.1. Sediment Type

Quantification of data from smear slides revealed that major constituents of the sediment core are clay minerals with dominant siliceous skeletons of microscopic marine organisms (radiolarians and diatoms) but where bioapatite fossil, zeolite, ferromanganese micronodules and feldspar could also be identified (Figure 3). The siliceous micro-organism fractions comprise, in general, more than 30% of the visual field, showing almost no change along the core. Accordingly, the sediment core was classified as a radiolarian clay.

4.2. Clay Minerals Compositions

Determination of the main clay minerals (montmorillonite, illite, chlorite, and kaolinite) of the core are shown in Table 1. Illite has the highest proportion (41.3–79.9%) with an average value of 59.5%, followed by montmorillonite (3.2–43.6%) with an average value of 20.2%. The contents of chlorite (8.2–17.0%) and kaolinite (5.3–9.7%) are relatively lower and their average values are 12.9% and 7.4%, respectively. The chemical index (CI) of illite was calculated from the ratio of peak areas at 5 Å and 10 Å from the XRD curves of glycol slices and the crystallinity of illite is represented by the peak Full Wave at Half Maximum (FWHM) at 10 Å. The range of illite CI is from 0.28 to 0.56 with a mean value of 0.47, and the crystallinity of illite was 0.16 ~0.30°Δ 2θ with a mean value of 0.24°Δ 2θ. The montmorillonite/illite (M/I) ratios vary from 0.04 to 1.05 and with a mean value of 0.39. The clay minerals assemblage composition of the sediment core is: illite–montmorillonite–chlorite–kaolinite.

4.3. Geochemical Characteristics of the Bulk Sediments

4.3.1. Major Elements

Overall, the SiO2 contents are the highest, ranging from 49.36% to 54.05% with an average value of 51.68%, followed by Al2O3 (14.43–12.55%) with an average value of 13.79%. P2O5, TiO2 and MnO contents are the lowest among the major elements, all of which are <1%. The detailed data are presented in Supplementary Table S1.
Figure 4 shows the trends of major element concentrations with increasing depth of the core. Among all trends, the contents of P2O5, MgO and CaO displayed the most significant and consistent increase with depth. In comparison, the increasing trends for Fe2O3 and Na2O seem much more moderate along the core, and Fe2O3, but not Na2O, decreases between 400–570 cm of the core. Opposite to P2O5, MgO and CaO, the trends for SiO2 and K2O gradually and consistently decrease from the surface to the bottom. The remaining oxides do not show a regular pattern of variation along the core.

4.3.2. Rare Earth Elements

The composition of the REY along the sediments is provided in Supplementary Table S2 and North American shale composite (NASC)-normalized REE patterns are shown in Figure 5. The total REY concentrations range between 321.9 ppm and 535.1 ppm, with a mean value of 454.7 ppm, which is significantly higher than that of NASC (177.4 ppm [32]). Light REE (LREE; La-Eu)/heavy REE (HREE; Gd-Lu) ratios range from 3.7 to 4.7 with an average value of 4.0. The values of cerium anomaly (δCe) and europium anomaly (δEu) are calculated by the following formulations respectively: δCe = 2CeN/LaN + PrN; and δEu = 2EuN/SmN + GdN, while the subscript N indicates the normalization by NASC value [32]. The δCe values of all samples are smaller than 1, with an average value of 0.59. Therefore, all samples show negative Ce anomalies (Figure 5). The δEu values range from 0.86 to 1.04 with an average value of 0.93, showing slightly negative Eu anomalies overall.
The REE distribution curves of all samples are quite similar (Figure 5), and are characterized by a loss of LREE and enrichments of HREE, obvious negative Ce anomalies, negative Nd anomalies, positive Sm anomalies, positive Gd anomalies, negative Dy anomalies and positive Ho anomalies.

4.4. Scanning Electron Microscopy

Scanning Electron Microscopy (SEM) analyses were conducted on the clay fraction of the layers between 180–190 cm from GC04 and compared with those of the GC1601 core (layer 30–40 cm). Figure 6 shows the microphotographs of clay minerals of these two cores. Specifically, Figure 6A to E refer to the micrographs of GC04, while F to I refer the micrographs of layer 30–40 cm of GC1601. Compared with the GC1601 (Figure 6h,i), the illite crystallization from GC04 is less developed (Figure 6c,d). Specifically, the former shows an obvious schistose structure, while the latter is separated from that of kaolinite under the electron microscope. In addition, illite and kaolinite of GC04 were often found grown together, forming an illite–kaolinite mixed layer (Figure 6d,e), which is in stark contrast with that of GC1601.

5. Discussion

5.1. Sediment Input Sources

5.1.1. Clay Minerals Evidence

Clay minerals are the most widely distributed minerals on the sea floor [33]. Their compositions and distribution patterns record information of transports, redepositions, and environmental evolutions, and can thus be used to clarify sedimentation and material source [34,35,36].
Illite, one of the most common clay minerals in marine sediments, is basically considered to be of terrestrial source, and is mainly imported into the ocean through rivers and wind [37,38]. The mean contents of illite in sediments of our study are 59.5% (Table 1), which is similar to that of Yellow River and Yangtze River sediments [39,40,41], in agreement with a terrestrial origin of illite. Small values of illite crystallinity indicate a better crystallization and refer to dry and cold climatic conditions, as illite chemical index (CI) is below 0.5, it usually indicates physical weathering [37]. The CI of illite ranges from 0.28 to 0.56, with a mean value of 0.47 (Table 1), and the illite crystallinity varies from 0.36 to 0.46°Δ 2θ, with a mean value of 0.42°Δ 2θ (Table 1), which are both very close to the CI and crystallinity of illite of similar sediment cores where a dominance of terrestrial material, such as East Asian dust, has been identified (e.g., GC1601 [37]). Considering that the study sediment was collected in an area far from the continent, a riverine input is improbable, we speculate that the high illite content in the study area mainly refers to eolian dust input.
There are two main sources for montmorillonite in marine sediments, terrestrial detritus, formed by hydrolysis of source rocks in continental warm to semi-arid climates, and submarine volcanic materials, which are formed by long-term marine hydrolysis and weathering [37]. The high content of montmorillonite generally indicates high volcanic hydrothermal activity [42,43]. When the volcanic materials in the sediments increase, the montmorillonites will increase accordingly [44], and their content can even reach more than 50% [37]. Similarly, as the sediment origin is mainly dominated by terrestrial material, the montmorillonite content would be relatively low, typically not exceeding 20%, while the illite content tends to be very high [45]. Therefore, montmorillonite always shows a negative correlation with illite, which is also well represented in this study (Figure 7).
The average content of montmorillonite in the sediments of this study can reach 20.2% and with the highest value up to 43.6% (Table 1), suggesting that the sediment material in the area is also likely to be affected by volcanic material input. Previous studies have also suggested that montmorillonite in this area may originate from the alteration of volcanic materials [24].
For a more visual representation of the source of clay minerals in sediments, a ternary diagram of clay mineral assemblages is shown in Figure 8. The majority of sediment samples in our study fall into the terrestrial source field, but some point to volcanic hydrothermal end-member, indicating that the clay mineral compositions of the core are characterized by the mixing of terrigenous materials and volcanic hydrothermal materials, and the supply of terrigenous materials is more obvious than latter.
Although the clay extraction process for GC04 and GC1601 is identical, the clay content extraction of GC04 is very low compared with GC1601. The clay minerals in GC1601 occupy more than 90% of the field of view [52], whereas the clay minerals in GC04 make up a very limited proportion of the overall field of view (Figure 3). Therefore, although illite content and clay mineral parameters (e.g., CI, illite crystallinity and M/I) are quite close, unlike GC1601, where the material source of sediment is clearly dominated by terrestrial sources [37,52], the dominant material source of GC04 may not be terrestrial. The GC04 is far away from the Asian continent so that the contribution of terrigenous aeolian dust should be relatively limited, with the result that illite that mainly derived from terrigenous aeolian dust accounts for a low proportion of total sediments. As mentioned earlier, since there is a relatively high content of montmorillonite in GC04, the core is likely to be affected by volcanic material input. At the same time, we also compared the microscopic morphology of clay minerals in our sediment sample with that of a core from the Northwest Pacific (GC1601; Figure 6). This comparison shows that the crystallization of illite phases in GC04 is lower than in GC1601 (Figure 6a,c,d,h,i). In GC04, there was also observed a larger amount of montmorillonite/illite mixed layers (Figure 6c–e), indicating that the montmorillonite would also have a volcanic hydrothermal origin. Considering the strong tectonic activity in the study area, combined with the result of the ternary diagram of the clay mineral assemblages, we propose that volcanic hydrothermal input is also considerable in the sediments of the study area.

5.1.2. REEs Evidence

REEs and their parameters act as powerful tools for indicating material sources of sediments [17], some parameters including Ce and Eu anomalies and fractionation of LREEs or middle REE (Sm-Dy; MREE) relative to HREEs, represent compositional differences between sediment origins, thus sediment origin can be successfully discriminated if the signals from source-related materials can be identified and distinguished [53,54].
The sources of marine sediments are complex, and include terrestrial (e.g., aeolian dust), biogenic (e.g., calcareous and siliceous biological fragments), authigenic (e.g., phillipsites) and local sources (e.g., volcanic ash and its altered materials [55]). In order to further discuss the potential sediment sources and their contribution in the study area, the REEs discriminant function (DF) was applied and the expression of DF is as follows:
DF = (E1X/E2X)/(E1L/E2L) − 1
where (E1X/E2X) represents the ratio of two REEs in sediments of the study area and (E1L/E2L) represents the ratio of two REEs of the geological body of possible sources. When the absolute value of DF is less than 0.5, it is considered that the two substances are similar, and the smaller the DF value, the higher the degree [52,56]. The chemical properties of the two elements should be as similar as possible and here we used La/Tb and La/ Yb as the discriminant REE ratios. The DF value of sediments in the study area and other geological bodies are calculated according to reported La/Tb and La/ Yb ratios from different sources, (Table 2).
The DF values of terrigenous, local, biogenic, and authigenic sources in the studied samples calculated using La/Tb and La/Yb are almost all less than 0.5, which indicates that the sediments in the area had input from these different sources. The DF value of diatom ooze is the lowest, ranging from 0.28 to 0.08, with a mean value of 0.18, which indicates that the compositions of GC04 are the closest to siliceous ooze, thus with a strong contribution of siliceous organisms. Smear slide identification also showed that siliceous organisms (including diatoms and radiolarians) occur in a high percentage (Figure 3), which is consistent with results showing that the excess silicon in sediment reaches more than 10%. TheGC04 core was collected at a water depth of 5144 m, which is deeper than the carbonate compensation depth (CCD), explaining the absence of calcareous organisms. An input of calcareous organisms to the sediments can thus be ignored when considering the influence of biogenic materials. The calculated DF values of terrigenous source, local source and authigenic sources present similar average values of 0.42, 0.48 and 0.45, respectively, revealing similar and consistent origins and inputs, with a little bit more input of terrigenous material, followed by the biogenic fraction, and finally by volcanic materials and their weathering products.
DF values of terrigenous sources and local sources confirm the inference that terrigenous sources may not be the dominant substance of the core and volcanic materials also make a certain contribution. The mean DF value of the authigenic source (seawater) suggests that sediments in the study area also received considerable material supply from these different sources. As for pelagic sediments, they are usually mixed with authigenic substances such as iron–manganese oxides, zeolites and bioapatite and smear slide identification also confirmed their obvious occurrence (Figure 3).
Previous studies have suggested that REE in deep-sea sediments derive directly from seawater [1,36], and the REE in GC04 gradually decrease from the surface to the bottom, indicating that the supply of substances in seawater representing authigenic sources to sediments gradually decreases. The P2O5, mainly in the form of authigenic apatite, also generally increases from the surface to the bottom (Figure 6), indicating that the supply of authigenic materials to the core decreases over time. Besides, the illite content of the core basically decrease gradually from the bottom to the surface (Table 1), which is somewhat similar to the profile change of Ti contents (Figure 6), which itself mainly represents a terrigenous source, while SiO2 also decreases gradually from the bottom to the surface overall, indicating that both the supply of terrigenous and biogenic materials increase during the deposition history of the core. Opposite to illite, an overall increase in montmorillonite from the surface to the bottom of GC04 (Table 1) suggests that the core was more strongly affected by volcanic hydrothermal materials (local sources) during the early deposition period.

5.2. Sedimentary Source Structure Properties

REE compositions can be affected by post-depositional diagenetic processes including changes in redox conditions and the formation of authigenic minerals, such as carbonates, phosphates, and Fe–Mn oxides [17,60]. Um et al. (2013) proposed that δCe gradually decreases in sediments from the continental shelf, continental slope sediments, and deep-sea basin, and the REE distribution patterns in sediments from different tectonic environment are significantly different [53]. Generally, the REE in terrestrial environments are enriched in LREEs as opposed to HREEs [56], while the pattern is opposite for seawater, deep-sea sediments, and marine authigenic biomass.
Here, in order to determine the source structure properties of sediments more directly in the study area, the NASC-normalized REE distribution patterns of rivers sediments, marginal sea sediments, back arc basins sediments, pelagic deep-sea basin sediments, marine authigenic biomass (biological phosphate) and seawater are compared (Figure 9). The interesting thing is that all these geological bodies exhibit different REE distribution curves, with river sediments, including those from the Yellow River, characterized with a distinct loss of HREE and enrichment of LREE (Figure 9a), markedly different from that of GC04. The REE distribution curve of marginal sea sediments, such as South China sediments, East China Sea sediments, and Japan Sea sediments, are similar to those of river sediments, but their right-tilted profile is not so obvious (Figure 9b). The REE distribution curves of sediments in the back arc basin (Philippine Basin) are characterized with slight HREE enrichment. By contrast, pelagic sediments are far away from the continent and receive little supply of terrigenous materials. Their REE distribution patterns (DSDP Site 170 and site 213) are similar to those of seawater (Figure 9d), with obvious negative Ce anomalies and relative enrichment of HREE, similar to that of GC04, indicating that the sedimentary source structure properties of sediments in the study area are mostly pelagic. The REE distribution pattern of bioapatite is highly consistent with that of seawater (Figure 9e), showing relative HREE enrichment characteristics and negative Ce anomalies, as compared with that of GC04 or pelagic sediments.
The location of GC04 is far away from the Asian continent and American continent and at a water depth >5000 m. It is thus expected that the sediments should suffer from a strong ‘seawater effect’. Such an effect can be interpreted as seawater being the direct source of some elements in sediments. For example, Ce3+ will be oxidized to Ce4+ via reaction on highly oxidative Mn–oxides, forming Ce(OH)4, and then co-precipitated into the Fe–Mn nodules as Fe–Mn oxides or hydroxides, resulting in the depletion of Ce in sediments and its enrichment in Fe-Mn nodules (Figure 9f) [28]. Authigenic materials such as bioapatites, which are highly enriched in REE, precipitate at the seafloor and influence the REE distribution patterns of sediments [18]. Generally, when the amount of REE provided by marine biomass is significantly higher than that from a terrestrial source, the distribution patterns of sediments will be more similar to that of seawater, which is a common characteristic of deep-sea pelagic sediments (Figure 9d).
As discussed above, Ce3+ in seawater tends to be oxidized to Ce4+, which results in significantly negative Ce anomalies in seawater, usually <1 [59]. Most of the marine authigenic materials inherit the REE distribution characteristics of seawater, also having low Ce values. Therefore, the δCe value can indicate the contribution of seawater as a source of materials to sediments to some extent. To facilitate the identification of the input sources of the study samples, data from sediments under different tectonic backgrounds are plotted into LREE/HREE vs. 1/δCe and LREE/HREE vs. Y/Ho diagram (Figure 10).
This plot shows a strong convergence between sediments or geological bodies from different tectonic environments, with sediments from the Japan Sea and South China Sea sediments located at the top of the diagrams, belonging to the sedimentary properties of the marginal sea. Our sediments fall between the marginal sea to pelagic transition sediment area and the seawater and marine authigenic area, which is the sedimentary properties of the pelagic sedimentary area (Figure 10a). In addition, Y/Ho is also used as an effective proxy of the influence of marine authigenic components [70]. The closer the Y/Ho value in sediment to seawater, the stronger the influence and the larger the proportion of components from seawater. For example, the Y/Ho ratios in seawater are close to 100 [59], while the Y/Ho ratio of bioapatite can be usually greater than 30. According to Figure 9b, the Japan Sea sediments and GC1601 fall into the weak seawater transformation area, while the study samples and other pelagic sediments basically fall in the weak area of strong seawater transformation, which further indicates that the sedimentary source structure properties of the GC04 are mostly pelagic.
The sedimentary model from the origin of the material in the studied sediments can be seen in Figure 11. B represents marginal sea sedimentation area located between the continents and open oceans. These areas such as the South China Sea are impacted by East Asian monsoons and westerly jets (Figure 11), receiving a large supply of terrigenous materials [71], and accordingly sediments in this area show evidence of terrigenous sources and have REE distribution profiles similar to those of terrigenous materials [72]. The C area represents back-arc deposition and is further away from the continent, has less contribution of terrigenous material, as compared with B. The D area, which is where our study samples were deposited, is even further away from the continent as compared with C (Figure 11), and represents the pelagic sedimentary area as in Figure 10.

6. Conclusions

(1)
The sediment type of the GC04 core is radiolarian clay, and the clay minerals assemblage composition of the sediment core is illite–montmorillonite–chlorite–kaolinite. Among the major elements of the core, SiO2 (49.36–54.05%) is the highest major component, followed by Al2O3 and Fe2O3. The REE distribution patterns of all samples are highly consistent, showing obvious negative Ce anomalies, slight Eu anomalies and with more obvious enrichment of HREE than of LREE.
(2)
The material sources of sediment in the study area are multi-source, including biogenic, terrigenous, authigenic and local sources, among which biological sources (marine biomass) contribute the most. Terrestrial, local and autochthonous sources also contribute, but to a significantly lesser extent than biogenic sources. The contributions of biogenic and terrigenous sources decrease from surface to bottom of GC04, which is opposite to that of authigenic and local sources.
(3)
According to the comparison between the characteristics of REE distribution patterns of sediments in the study area and other tectonic environments, as well as combined with the discrimination of relevant REE parameters, it is suggested that the sedimentary source structure of the study area refer to pelagic sedimentary properties which suffer from a strong ‘seawater effect’.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jmse10081052/s1, Table S1: Major elements compositions of GC04; Table S2: Rare earth elements compositions of GC04.

Author Contributions

Conceptualization, W.M., C.T. and Z.Q.; methodology, Z.Q.; software, Z.Q. and S.H.; validation, W.M. and C.T.; formal analysis, Z.Q.; investigation, W.M., C.T. and Z.Q.; resources, W.M., C.T. and Á.A.D.; data curation, Z.Q. and W.M.; writing—original draft preparation, Z.Q.; writing—review and editing, all authors; visualization, W.M., C.T., Á.A.D., S.H., Y.S., K.Y. and W.Z.; supervision, W.M. and C.T.; project administration, W.M., C.T. and Á.A.D.; funding acquisition, W.M., C.T. and Á.A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was jointly funded by National Natural Science Foundation of China (Grant No. 42127807), Key Research and Development Program of Zhejiang Province (Grant No. 2021C03016), the China Ocean Mineral Resources R&D Association (COMRA) Project under contract no. DY135-C1–1-05, Macao Science and Technology Development Fund (FDCT) project 0041/2021/A1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

We thank Fang Y for assisting SEM-EDS analysis and Zhu JH for assisting XRF and ICP-MS analysis, and we also thank all the members in DY-40B cruise for their hard work in sampling process.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yasukawa, K.; Nakamura, K.; Fujinaga, K.; Iwamori, H.; Kato, Y. Tracking the spatiotemporal variations of statistically independent components involving enrichment of rare-earth elements in deep-sea sediments. Sci. Rep. 2016, 6, 29603. [Google Scholar] [CrossRef] [PubMed]
  2. Dias, S.; Mills, R.; Taylor, R.; Ferreira, P.; Barriga, F. Geochemistry of a sediment push-core from the Lucky Strike hydrothermal field, Mid-Atlantic Ridge. Chem. Geol. 2008, 247, 339–351. [Google Scholar] [CrossRef]
  3. Yasukawa, K.; Ohta, J.; Miyazaki, T.; Vaglarov, B.S.; Chang, Q.; Ueki, K.; Toyama, C.; Kimura, J.I.; Tanaka, E.; Nakamura, K. Statistic and Isotopic Characterization of Deep-Sea Sediments in the Western North Pacific Ocean: Implications for Genesis of the Sediment Extremely Enriched in Rare Earth Elements. Geochem. Geophy. Geosy. 2019, 20, 3402–3430. [Google Scholar] [CrossRef]
  4. Dunlea, A.G.; Murray, R.W.; Sauvage, J.; Pockalny, R.A.; Spivack, A.J.; Harris, R.N.; D Hondt, S. Cobalt-based age models of pelagic clay in the South Pacific Gyre. Geochem. Geophy. Geosy. 2015, 16, 2694–2710. [Google Scholar] [CrossRef]
  5. Halbach, P.; Hebisch, U.; Scherhag, C. Geochemical variations of ferromanganese nodules and crusts from different provinces of the Pacific Ocean and their genetic control. Chem. Geol. 1981, 34, 3–17. [Google Scholar] [CrossRef]
  6. Dymond, J.; Lyle, M.; Finney, B.; Piper, D.Z.; Murphy, K.; Conard, R.; Pisias, N. Ferromanganese nodules from MANOP Sites H, S, and R—Control of mineralogical and chemical composition by multiple accretionary processes. Geochim. Cosmochim. Acta 1984, 48, 931–949. [Google Scholar] [CrossRef]
  7. Koschinsky, A.; Hein, J.R.; Kraemer, D.; Foster, A.L.; Halbach, P. Platinum enrichment and phase associations in marine ferromanganese crusts and nodules based on a multi-method approach. Chem. Geol. 2019, 539, 119426. [Google Scholar] [CrossRef]
  8. Qiu, Z.; Dong, Y.; Ma, W.; Zhang, W.; KY, Y.; Zhao, H. Geochemical characteristics of platinum-group elements in polymetallic nodules from the Northwest Pacific Ocean. Acta Oceanol. Sin. 2020, 39, 34–42. [Google Scholar] [CrossRef]
  9. Menendez, A.; Rachael, H.J.; Anna, L.; Douglas, C.; Kate, P. Controls on the chemical composition of ferromanganese nodules in the Clarion-Clipperton Fracture Zone, eastern equatorial Pacific. Mar. Geol. 2019, 409, 1–14. [Google Scholar] [CrossRef]
  10. Wegorzewski, A.V.; Kuhn, T. The influence of suboxic diagenesis on the formation of manganese nodules in the Clarion Clipperton nodule belt of the Pacific Ocean. Mar. Geol. 2014, 357, 123–138. [Google Scholar] [CrossRef]
  11. Koschinsky, A.; Bau, M.; Marbler, H.; Schmidt, K. Rare valuable metals in marine ferromanganese nodules–contents and accumulation processes. Z. Für Angew. Geologie. 2010, 51, 33–39. [Google Scholar]
  12. Heller, C.; Kuhn, T.; Versteegh, G.J.; Wegorzewski, A.V.; Kasten, S. The geochemical behavior of metals during early diagenetic alteration of buried manganese nodules. Deep Sea Res. Part I 2018, 142, 16–33. [Google Scholar] [CrossRef]
  13. Guan, Y.; Sun, X.; Ren, Y.; Jiang, X. Mineralogy, geochemistry and genesis of the polymetallic crusts and nodules from the South China Sea. Ore. Geol. Rev. 2017, 89, 206–227. [Google Scholar] [CrossRef]
  14. Cha, H.J.; Man, S.C.; Lee, C.B.; Shin, D.H. Geochemistry of surface sediments in the southwestern East/Japan Sea. J. Asian Earth Sci. 2007, 29, 685–697. [Google Scholar] [CrossRef]
  15. Dou, Y.; Yang, S.; Liu, Z.; Clift, P.D.; Shi, X.; Hua, Y.; Berne, S. Provenance discrimination of siliciclastic sediments in the middle Okinawa Trough since 30 ka: Constraints from rare earth element compositions. Mar. Geol. 2010, 275, 212–220. [Google Scholar] [CrossRef]
  16. Cha, H.; Lee, C.; Kim, B.; Choi, M.; Ruttenberg, K. Early diagenetic redistribution and burial of phosphorus in the sediments of the southwestern East Sea (Japan Sea). Mar. Geol. 2005, 216, 127–143. [Google Scholar] [CrossRef]
  17. Song, Y.H.; Man, S.C. REE geochemistry of fine-grained sediments from major rivers around the Yellow Sea. Chem. Geol. 2009, 26, 328–342. [Google Scholar] [CrossRef]
  18. Qiu, Z.; Ma, W.; Zhang, X.; Dong, Y.; Zhang, W.; Yang, K. Geochemical characteristics of surface sediments in the Northwest Pacific Ocean and their source-indicating significance. J. Zhejiang Univ. Sci. Ed. 2020, 47, 345–354. (In Chinese) [Google Scholar]
  19. Munksgaard, N.C.; Lim, K.; Parry, D.L. Rare earth elements as provenance indicators in North Australian estuarine and coastal marine sediments. Estuar. Coast Shelf Sci. 2003, 57, 399–409. [Google Scholar] [CrossRef]
  20. Fangjian, X.; Anchun, L.; Tiegang, L.; Kehui, X.U.; Shiyue, C.; Longwei, Q.; Yingchang, C. Rare earth element geochemistry in the inner shelf of the East China Sea and its implication to sediment provenances. J. Rare Earth. 2011, 2929, 702–709. [Google Scholar]
  21. Nie, Y.; Liu, X.; Emslie, S.D. Distribution and sources of rare earth elements in ornithogenic sediments from the Ross Sea region, Antarctica. Microchem. J. 2014, 114, 247–260. [Google Scholar] [CrossRef]
  22. Reykhard, L.Y.; Shulga, N.A. Fe-Mn nodule morphotypes from the NE Clarion-Clipperton Fracture Zone, Pacific Ocean: Comparison of mineralogy, geochemistry and genesis. Oro. Geol. Rev. 2019, 110, 102933. [Google Scholar] [CrossRef]
  23. Liu, J.; Shi, X.; Chen, L.; Huang, Y.; Wang, Y.; Cui, Y.; Bu, W. REEs and εNd of clay fractions in sediments from the eastern Pacific Ocean—Evidence for clay sources. Sci. China Ser. D Earth Sci. 2004, 34, 552–561. (In Chinese) [Google Scholar]
  24. Fu, F.; Zheng, Y.; Yao, X.Y.; Ni, J.Y. Mineral and geochemical characteristics of surface sediment clays in the western part of the eastern Pacific CC area. J. Mar. Sci. 2017, 3535, 55–65. (In Chinese) [Google Scholar]
  25. Jiang, L.; Wang, W.; Liang, J.; Dai, S.; Jiang, M.; Tao, W.; Ma, X. Geochemical Characteristics and Genetic Mechanisms of Sediments from Western Clarion-Clipperton Zone, Pacific Ocean. Acta Sedimentol. Sin. 2021, 39, 1185–1197. (In Chinese) [Google Scholar]
  26. Ni, J.; Zhou, H.; Pan, J.; Zhao, H.; Hu, C.; Wang, F. Geochemical characteristics of sediments from the COMRA registered pioneer area (CRPA), equatorial northeastern Pacific Ocean. Acta Oceanol Sin. 2001, 4, 553–561. [Google Scholar]
  27. Hein, J.R.; Mizell, K.; Koschinsky, A.; Conrad, T.A. Deep-ocean mineral deposits as a source of critical metals for high- and green-technology applications: Comparison with land-based resources. Oro. Geol. Rev. 2013, 51, 1–14. [Google Scholar] [CrossRef]
  28. Qiu, Z.; Ma, W.; Yang, K. Characteristics and significance of platinum-group elements (PGEs) in polymetallic nodules from CCZ, East Pacific. Geol. Sci. Technol. Inf. 2020, 39, 109–117. (In Chinese) [Google Scholar]
  29. Müller, R.D.; Sdrolias, M.; Gaina, C.; Roest, W.R. Age, spreading rates, and spreading asymmetry of the World’s Ocean crust. Geochem. Geophy. Geosy. 2008, 9, Q04006. [Google Scholar] [CrossRef]
  30. Jeong, K.S.; Kang, J.K.; Chough, S.K. Sedimentary processes and manganese nodule formation in the Korea Deep Ocean Study (KODOS) area, western part of Clarion-Clipperton fracture zones, northeast equatorial Pacific. Mar. Geol. 1994, 122, 125–150. [Google Scholar] [CrossRef]
  31. Biscaye, P.E. Mineralogy and sedimentation of recent deep-sea clay in atlantic ocean and adjacent areas and oceans. Geol. Soc. Am. Bull. 1965, 76, 803–832. [Google Scholar] [CrossRef]
  32. Haskin, L.A.; Haskin, M.A.; Frey, F.A.; Wildeman, T.R. Relative and Absolute Terrestrial Abundances of the Rare Earths. Orig. Distrib. Elem. Symp. 1968, 1, 889–912. [Google Scholar]
  33. Sattarova, V.V.; Artemova, A.V. Geochemical and micropaleontological character of Deep-Sea sediments from the Northwestern Pacific near the Kuril–Kamchatka Trench. Deep Sea Res. Part II 2015, 111, 10–18. [Google Scholar] [CrossRef]
  34. Khan, M.; Liu, J.; Liu, S.; Seddique, A.A.; Cao, L.; Rahman, A. Clay mineral compositions in surface sediments of the Ganges-Brahmaputra-Meghna River system of Bengal Basin, Bangladesh. Mar. Geol. 2019, 412, 27–36. [Google Scholar] [CrossRef]
  35. Ehrmann, W. Implications of late Eocene to early Miocene clay mineral assemblages in McMurdo Sound (Ross Sea, Antarctica) on paleoclimate and ice dynamics. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1998, 139, 213–231. [Google Scholar] [CrossRef]
  36. Yasukawa, K.; Liu, H.; Fujinaga, K.; Machida, S.; Haraguchi, S.; Ishii, T.; Nakamura, K.; Kato, Y. Geochemistry and mineralogy of REY-rich mud in the eastern Indian Ocean. J. Asian Earth Sci. 2014, 93, 25–36. [Google Scholar] [CrossRef]
  37. Qiu, Z.; Ma, W.; Tao, C.; Koschinsky, A.; Hu, S. Clay Minerals and Sr-Nd Isotope Compositions of Core CG 1601 in the Northwest Pacific: Implications for Material Source and Rare Earth Elements Enrichments. Minerals 2022, 12, 287. [Google Scholar] [CrossRef]
  38. Saburo, A.; Toshio, S. Clay minerals in the deep-sea cores from the North Pacific. J. Oceanogr. 1972, 28, 138–144. [Google Scholar]
  39. Ren, M.; Shi, Y. Sediment discharge of the Yellow River (China) and its effect on the sedimentation of the Bohai and the Yellow Sea. Cont. Shelf Res. 1986, 6, 785–810. [Google Scholar] [CrossRef]
  40. Yang, S.Y.; Jung, H.S.; Lim, D.I.; Li, C.X. A review on the provenance discrimination of sediments in the Yellow Sea. Earth-Sci. Rev. 2003, 63, 93–120. [Google Scholar] [CrossRef]
  41. Yang, S.Y.; Jung, H.S.; Choi, M.S.; Li, C.X. The rare earth element compositions of the Changjiang (Yangtze) and Huanghe (Yellow) river sediments. Earth Planet Sci. Lett. 2002, 201, 407–419. [Google Scholar] [CrossRef]
  42. Köhler, B.; Singer, A.; Stoffers, P. Biogenic nontronite from marine white smoker chimneys. Clays Clay Miner. 1994, 42, 689–701. [Google Scholar] [CrossRef]
  43. Corliss, J.B.; Lyle, M.; Dymond, J.; Crane, K. The chemistry of hydrothermal mounds near the Galapagos Rift. Earth Planet Sci. Lett. 1978, 40, 12–24. [Google Scholar] [CrossRef]
  44. Gingele, F.X. Holocene climatic optimum in Southwest Africa—Evidence from the marine clay mineral record. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1996, 122, 77–87. [Google Scholar] [CrossRef]
  45. Wan, S.; Yu, Z.; Clift, P.D.; Sun, H.; Li, A.; Li, T. History of Asian eolian input to the West Philippine Sea over the last one million years. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2012, 326, 152–159. [Google Scholar] [CrossRef]
  46. Shi, Y.X.; Dai, X.R.; Li, J.T.; Wang, J.C. Clay minerals in the Late Interglacial Lanzhou Loess Record and their environmental significance. Mar. Geol. Quat. Geol. 1997, 17, 87–94. (In Chinese) [Google Scholar]
  47. Wan, S.; Li, A.; Clift, P.D.; Stuut, J.W. Development of the East Asian monsoon: Mineralogical and sedimentologic records in the northern South China Sea since 20 Ma. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2007, 254, 561–582. [Google Scholar] [CrossRef]
  48. Zhang, D. Clay minerals composition and distribution in the Mariana Trough. Adv. Mar. Sci. 1994, 12, 32–39. [Google Scholar]
  49. Chi, Y.; Li, A.; Jiang, F.; Wang, M. Characterization of clay mineral assemblages and physical source analysis in the eastern Luzon Sea. Mar. Sci. 2009, 33, 80–88. (In Chinese) [Google Scholar]
  50. Shi, X.F.; Chen, L.R.; Li, K.Y.; Wang, Z.L. Genetic mineralogy of clay sediments in the western Philippine Sea. Mar. Geol. Quat. Geol. 1995, 15, 61–72. (In Chinese) [Google Scholar]
  51. Liu, Z.; Zhao, Y.; Colin, C.; Siringan, F.P.; Wu, Q. Chemical weathering in Luzon, Philippines from clay mineralogy and major-element geochemistry of river sediments. Appl. Geochem. 2009, 24, 2195–2205. [Google Scholar] [CrossRef]
  52. Qiu, Z.; Ma, W.; Zhang, X.; Ren, J. Geochemical characteristics of sediments in the southeastern waters of Nanjiao Island and their significance in indicating rare earth resources. Geol. Sci. Technol. Inf. 2019, 38, 205–214. (In Chinese) [Google Scholar]
  53. Um, I.K.; Man, S.C.; Bahk, J.J.; Song, Y.H. Discrimination of sediment provenance using rare earth elements in the Ulleung Basin, East/Japan Sea. Mar. Geol. 2013, 346, 208–219. [Google Scholar] [CrossRef]
  54. Webb, G.E.; Kamber, B.S. Rare earth elements in Holocene reefal microbialites: A new shallow seawater proxy. Geochim. Cosmochim. Acta 2000, 64, 1557–1565. [Google Scholar] [CrossRef]
  55. Shepard, F.P.; Moore, D.G. Post-glacial stratigraphy alog the central Texas coast. J. Sediment Res. 1954, 24, 140. [Google Scholar] [CrossRef]
  56. Jung, H.S.; Lim, D.; Choi, J.Y.; Yoo, H.S.; Rho, K.C.; Lee, H.B. Rare earth element compositions of core sediments from the shelf of the South Sea, Korea: Their controls and origins. Cont. Shelf Res. 2012, 48, 75–86. [Google Scholar] [CrossRef]
  57. Chen, L.; Zhang, K.; Fu, J.; Chen, H.; Li, X.; Li, Z. Geochemical characteristics of trace and rare earth elements and their material source indication significance since the deposition of S_2 in the loess paleosol of Mang Shan. Acta Sci. Nat. Univ. Sunyatseni 2018, 57, 14–23. (In Chinese) [Google Scholar]
  58. Davis, A.S.; Gray, L.B.; Clague, D.A.; Hein, J.R. The Line Islands revisited: New 40Ar/39Ar geochronologic evidence for episodes of volcanism due to lithospheric extension. Geochem. Geophys. Geosyst. 2013, 3, 1–28. [Google Scholar] [CrossRef]
  59. Alibo, D.S.; Nozaki, Y. Rare earth elements in seawater: Particle association, shale-normalization, and Ce oxidation. Geochim. Cosmochim. Acta 1999, 63, 363–372. [Google Scholar] [CrossRef]
  60. Haley, B.A.; Klinkhammer, G.P.; Mcmanus, J. Rare earth elements in pore waters of marine sediments. Geochim. Cosmochim. Acta 2004, 68, 1265–1279. [Google Scholar] [CrossRef]
  61. Gaillardet, J.; Dupré, B.; Allègre, C.J. Geochemistry of large river suspended sediments: Silicate weathering or recycling tracer? Geochim. Cosmochim. Acta 1999, 63, 4037–4051. [Google Scholar] [CrossRef]
  62. Zhang, C.; Wang, L. Multi-element geochemistry of sediments from the Pearl River system, China. Appl. Geochem. 2001, 16, 1251–1259. [Google Scholar] [CrossRef]
  63. Xu, F.; Hu, B.; Dou, Y.; Liu, X.; Wan, S.; Xu, Z.; Tian, X.; Liu, Z.; Yin, X.; Li, A. Sediment provenance and paleoenvironmental changes in the northwestern shelf mud area of the South China Sea since the mid-Holocene. Cont. Shelf Res. 2017, 144, 21–30. [Google Scholar] [CrossRef]
  64. Otosaka, S.; Togawa, O.; Baba, M.; Karasev, E.; Volkov, Y.N.; Omata, N.; Noriki, S. Lithogenic flux in the Japan Sea measured with sediment traps. Mar. Geol. 2004, 91, 143–163. [Google Scholar] [CrossRef]
  65. Chu, Z.; Hu, Q.; Liu, J.; Gao, J.; Zhu, A.; Gao, J.; Zhang, H. Geochemical characteristics of rare earth elements in surface sediments of the West Philippine Sea and the significance of source indication. Mar. Geol. Quat. Geol. 2016, 36, 53–62. (In Chinese) [Google Scholar]
  66. Kato, Y.; Fujinaga, K.; Nakamura, K.; Takaya, Y.; Kitamura, K.; Ohta, J.; Toda, R.; Nakashima, T.; Iwamori, H. Deep-sea mud in the Pacific Ocean as a potential resource for rare-earth elements. Nat. Geosci. 2011, 4, 535–539. [Google Scholar] [CrossRef]
  67. Kon, Y.; Hoshino, M.; Sanematsu, K.; Morita, S.; Tsunematsu, M.; Okamoto, N.; Yano, N.; Tanaka, M.; Takagi, T. Geochemical Characteristics of Apatite in Heavy REE-rich Deep-Sea Mud from Minami-Torishima Area, Southeastern Japan. Resour. Geol. 2014, 64, 47–57. [Google Scholar] [CrossRef]
  68. Wang, F.; He, G.; Sun, X.; Yang, Y.; Zhao, T. Study of rare earth element fugacity carriers in Pacific rare earth-rich deep-sea sediments. Acta Petrol. Sin. 2016, 32, 2057–2068. (In Chinese) [Google Scholar]
  69. Zhang, X.; Tao, C.; Shi, X.; Huaiming, L.I.; Huang, M.U.; Huang, D. Geochemical characteristics of REY-rich pelagic sediments from the GC02 in central Indian Ocean Basin. J. Rare Earths 2017, 35, 1047–1058. [Google Scholar] [CrossRef]
  70. Johannesson, K.H. “Rare earth element geochemistry of scleractinian coral skeleton during meteoric diagenesis: A sequence through”. Sedimentology 2011, 59, 729–732. [Google Scholar] [CrossRef]
  71. Liu, J.; Xiang, R.; Chen, Z.; Chen, M.; Yan, W.; Zhang, L.; Chen, H. Sources, transport and deposition of surface sediments from the South China Sea. Deep Sea Res. Part I 2013, 71, 92–102. [Google Scholar] [CrossRef]
  72. Cwab, C.; Lza, C.; Ylab, C.; Jha, C.; Hca, C.; Ming, S. Geochemical record of the sediments in the continental shelf of the northwestern South China Sea: Implications for the provenance and sedimentary evolution. Mar. Geol. 2021, 440, 106582. [Google Scholar]
Figure 1. Core sampling location of GC04 in western Clarion Fracture Zone and other reference samples (GC1601, DSDP Site 170, DSDP Site 163). Gray arrow indicates directions of the westerly jet stream associated with the migration of terrigenous eolian dust. The map was created using GeoMapApp.
Figure 1. Core sampling location of GC04 in western Clarion Fracture Zone and other reference samples (GC1601, DSDP Site 170, DSDP Site 163). Gray arrow indicates directions of the westerly jet stream associated with the migration of terrigenous eolian dust. The map was created using GeoMapApp.
Jmse 10 01052 g001
Figure 2. Photographs of GC04 core. The black inverted triangle represents the sampling location for clay analysis.
Figure 2. Photographs of GC04 core. The black inverted triangle represents the sampling location for clay analysis.
Jmse 10 01052 g002
Figure 3. The optical photomicrographs of samples from the core. Abbreviations: Rad, Radiolaria; Dia, di-atom; Bia, bioapatite fossil; Zeo, zeolite; Min, micronodule; Fel, feldspar. (ai) represent the composition of substances under the microscope at different depths of GC04 respectively.
Figure 3. The optical photomicrographs of samples from the core. Abbreviations: Rad, Radiolaria; Dia, di-atom; Bia, bioapatite fossil; Zeo, zeolite; Min, micronodule; Fel, feldspar. (ai) represent the composition of substances under the microscope at different depths of GC04 respectively.
Jmse 10 01052 g003
Figure 4. Profile distribution of major elements along the core.
Figure 4. Profile distribution of major elements along the core.
Jmse 10 01052 g004
Figure 5. The NASC-normalized REE patterns in GC04 core. Note: δEu value of layer 515–520 cm is just 0.55, significantly lower than the adjacent layers, which may have been caused by systematic error, therefore, this group of data is meaningless and thus we will not emphasize a discussion of these data. NASC-normalized values are given by [32].
Figure 5. The NASC-normalized REE patterns in GC04 core. Note: δEu value of layer 515–520 cm is just 0.55, significantly lower than the adjacent layers, which may have been caused by systematic error, therefore, this group of data is meaningless and thus we will not emphasize a discussion of these data. NASC-normalized values are given by [32].
Jmse 10 01052 g005
Figure 6. SEM images of clay minerals of 1ayer 180–190 cm of GC04 (ae) and layer 30–40 cm of GC1601 (fi). (a,b) large scale microphotographs of clay minerals; (ce) montmorillonite illite layer; (f,g) large scale microphotographs of clay minerals; (h,i) schistose illite with good crystalline morphology.
Figure 6. SEM images of clay minerals of 1ayer 180–190 cm of GC04 (ae) and layer 30–40 cm of GC1601 (fi). (a,b) large scale microphotographs of clay minerals; (ce) montmorillonite illite layer; (f,g) large scale microphotographs of clay minerals; (h,i) schistose illite with good crystalline morphology.
Jmse 10 01052 g006
Figure 7. Correlation between montmorillonite and illite in GC04 and GC1601, (a) montmorillonite–illite, the area highlighted by the black dotted rectangle with the letter b is shown in (b).
Figure 7. Correlation between montmorillonite and illite in GC04 and GC1601, (a) montmorillonite–illite, the area highlighted by the black dotted rectangle with the letter b is shown in (b).
Jmse 10 01052 g007
Figure 8. Ternary diagram showing variation in clay mineral compositions of sediments from our study and other geological regions. CG1601 sediment data are from [37], Chinese loess data are from [46], Yellow River sediments data are from [39,40,41], Yangtze River sediments data are from [47], Mariana Trough sediments data are from [48], Philippine Trench sediments data are from [49,50], Luzon Island sediments data are from [51].
Figure 8. Ternary diagram showing variation in clay mineral compositions of sediments from our study and other geological regions. CG1601 sediment data are from [37], Chinese loess data are from [46], Yellow River sediments data are from [39,40,41], Yangtze River sediments data are from [47], Mariana Trough sediments data are from [48], Philippine Trench sediments data are from [49,50], Luzon Island sediments data are from [51].
Jmse 10 01052 g008
Figure 9. The NASC-normalized REE patterns in our study and comparison with other geological bodies, all the REE values of geological bodies are averaged. (a) Yangtze River sediments data are from [61], Yellow River sediments data are from [61], Pearl River sediments data are from [62], Mekong River sediments data are from [61]; (b) South China Sea sediments data are from [63], East China Sea sediments data are from [20], Japan Sea sediments data are from [64]; (c) Philippine Basin sediments data are from [65]; (d) Pacific pelagic sediments data are from [66], Indian Ocean deep-sea sediments data (Site 213) are from [36]; (e) the seawater values are given by [59], bioapatite data are from [67]; (f) Polymetallic nodules data in CCZ are from [28], the black line and red line represent the average value of nodules and GC04 respectively.
Figure 9. The NASC-normalized REE patterns in our study and comparison with other geological bodies, all the REE values of geological bodies are averaged. (a) Yangtze River sediments data are from [61], Yellow River sediments data are from [61], Pearl River sediments data are from [62], Mekong River sediments data are from [61]; (b) South China Sea sediments data are from [63], East China Sea sediments data are from [20], Japan Sea sediments data are from [64]; (c) Philippine Basin sediments data are from [65]; (d) Pacific pelagic sediments data are from [66], Indian Ocean deep-sea sediments data (Site 213) are from [36]; (e) the seawater values are given by [59], bioapatite data are from [67]; (f) Polymetallic nodules data in CCZ are from [28], the black line and red line represent the average value of nodules and GC04 respectively.
Jmse 10 01052 g009
Figure 10. (a) LREE/HREE vs. 1/δCe and (b) LREE/HREE vs. Y/Ho diagrams. The data are taken from: bioapatite-1 [67]; bioapatite-2 [68]; DSDP 163 [66]; Indian Ocean deep-sea sediments values [69]; Diatom ooze values [36]; Core CC1601 [37]. Other data sources are the same as in Figure 8.
Figure 10. (a) LREE/HREE vs. 1/δCe and (b) LREE/HREE vs. Y/Ho diagrams. The data are taken from: bioapatite-1 [67]; bioapatite-2 [68]; DSDP 163 [66]; Indian Ocean deep-sea sediments values [69]; Diatom ooze values [36]; Core CC1601 [37]. Other data sources are the same as in Figure 8.
Jmse 10 01052 g010
Figure 11. Sedimentary profile from continent to the ocean. A represents inland sedimentation (inland rivers or lakes), B represents marginal sea sedimentation, C represents back-arc basin sedimentation, D represents pelagic sedimentation.
Figure 11. Sedimentary profile from continent to the ocean. A represents inland sedimentation (inland rivers or lakes), B represents marginal sea sedimentation, C represents back-arc basin sedimentation, D represents pelagic sedimentation.
Jmse 10 01052 g011
Table 1. Clay minerals compositions of GC04.
Table 1. Clay minerals compositions of GC04.
Layer/cmRelative Content of Four Clay Minerals (%)M/IIllite
MontmorilloniteIlliteKaoliniteChloriteCICrystallinity
0–1015.060.97.117.00.250.280.222
60–7010.868.07.014.20.160.450.199
180–19017.957.89.714.50.310.430.297
240–25017.861.56.713.90.290.440.248
270–28025.450.99.514.10.500.440.283
300–31015.463.87.113.70.240.440.194
390–4003.274.47.914.50.040.500.161
420–43035.250.56.18.20.700.440.286
450–4605.979.95.98.30.070.500.245
480–49023.354.68.713.30.430.540.225
510–52013.562.99.314.30.210.560.246
540–55035.546.65.312.60.760.500.225
560–57043.641.36.38.81.050.550.252
Max43.679.99.717.01.050.560.297
Min3.241.35.38.20.040.280.161
Average20.259.57.412.90.390.470.237
Note: M/I-montmorillonite/illite; CI-chemical index.
Table 2. The DF values of sediments of our study.
Table 2. The DF values of sediments of our study.
RatiosTerrigenous SourceLocal SourceBiogenic SourcesAuthigenic Source
NASCChinese LoessSeamount BasaltDiatom OozeSeawater
La/Yb0.330.380.540.280.49
La/Tb0.460.460.410.080.41
Average0.400.420.480.180.45
Data source: Chinese loess [57]; Seamount basalt [58]; Diatom ooze [36]; Seawater [59].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qiu, Z.; Tao, C.; Ma, W.; Dias, Á.A.; Hu, S.; Shao, Y.; Yang, K.; Zhang, W. Material Source of Sediments from West Clarion–Clipperton Zone (Pacific): Evidence from Rare Earth Element Geochemistry and Clay Minerals Compositions. J. Mar. Sci. Eng. 2022, 10, 1052. https://doi.org/10.3390/jmse10081052

AMA Style

Qiu Z, Tao C, Ma W, Dias ÁA, Hu S, Shao Y, Yang K, Zhang W. Material Source of Sediments from West Clarion–Clipperton Zone (Pacific): Evidence from Rare Earth Element Geochemistry and Clay Minerals Compositions. Journal of Marine Science and Engineering. 2022; 10(8):1052. https://doi.org/10.3390/jmse10081052

Chicago/Turabian Style

Qiu, Zhongrong, Chunhui Tao, Weilin Ma, Ágata Alveirinho Dias, Siyi Hu, Yuexiao Shao, Kehong Yang, and Weiyan Zhang. 2022. "Material Source of Sediments from West Clarion–Clipperton Zone (Pacific): Evidence from Rare Earth Element Geochemistry and Clay Minerals Compositions" Journal of Marine Science and Engineering 10, no. 8: 1052. https://doi.org/10.3390/jmse10081052

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