Next Article in Journal
Study on Aeolian Sand-Box Backfill Geomechanical Characteristics and Overlying Strata Control Effects
Previous Article in Journal
Permian Granitic Plutons from the Northern Margin of the North China Craton: Implications for the Tectonic Evolution of the Central Asian Orogenic Belt
Previous Article in Special Issue
New Approach for Direct Determination of Manganese Valence State in Ferromanganese Nodules by X-ray Fluorescence Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Semi-Quantitative Analyses of Ferromanganese Nodules from the Pacific Ocean Using Synchrotron X-ray Powder Diffraction

1
Ocean Georesources Research Department, Korea Institute of Ocean Science & Technology, Busan 49111, Republic of Korea
2
Department of Geological Sciences, Pusan National University, Busan 46241, Republic of Korea
3
Department of Oceanography, Pusan National University, Busan 46241, Republic of Korea
4
Department of Earth and Environmental Sciences, Chonnam National University, Gwangju 61186, Republic of Korea
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(12), 1555; https://doi.org/10.3390/min13121555
Submission received: 27 November 2023 / Revised: 14 December 2023 / Accepted: 15 December 2023 / Published: 18 December 2023

Abstract

:
The geochemistry and mineralogy of Mn nodules offer crucial insights into the origins, environmental changes, and distribution of abyssal resources. However, the conventional laboratory X-ray diffractometer, usually employed for semi-quantitative analysis of mineral composition in Mn nodules, often fails to sufficiently detect minor phases due to beam flux limitations and high background signals. In this study, we investigated differences in manganate composition, even when comprising around 1% of the phase fraction, in two manganese nodules (KC-8 and KODOS-10) using high-resolution synchrotron X-ray diffraction. The Mn/Fe ratios of KC-8 and KODOS-10 were 1.32 and 6.24, respectively, indicating that KC-8 and KODOS-10 were predominantly formed in hydrogenetic and diagenetic environments. Both samples contained quartz, vernadite, buserite, and feldspar. Todorokite and illite were exclusively observed in KODOS-10. In KC-8, the phase fractions of vernadite and buserite among manganates ranged from 94(5)%–100(4)% and 6(1)%–0%, respectively. However, in KODOS-10, the fractions of vernadite, buserite, and todorokite ranged from 47(1)%–56(2)%, 33.6(4)%–40.1(3)%, and 10(3)%–16.3(8)%, respectively.

1. Introduction

Manganese (Mn) nodules, which are primarily composed of ferromanganese oxides, are metal-contained minerals found in deep-sea environments. These nodules have received significant economic interest owing to the abundance of metals contained within such as Ni, Cu, and Co and the recent increase in demand for secondary batteries as a component of electric vehicles [1,2,3]. They exhibit widespread distribution across the Pacific and Indian Oceans, alongside various marine environments globally, including the Atlantic and Arctic Oceans, as well as the Baltic Sea [4,5,6,7,8,9,10,11,12,13,14,15]. Previous studies on Mn nodules have mainly focused on their morphological, structural, and geochemical characteristics [7,16,17], and mineral composition. However, minor ferromanganese oxides are often inadequately observed due to limitations of conventional laboratory X-ray diffractometers (LXRD), including restricted beam flux, suboptimal resolution, and low signal-to-background ratio.
This study aimed to comprehensively analyze manganese oxide minerals within Mn nodules obtained from the Magellan Seamount in the southwestern Pacific Ocean. For this purpose, synchrotron-based X-ray diffraction (SXRD) at the Pohang Accelerator Laboratory (PAL) was employed. The objective was to achieve a semi-quantitative and detailed understanding of mineral compositions.
Synchrotron X-ray diffraction (SXRD) stands out in providing highly precise information about crystallized materials. While its setup resembles that of laboratory X-ray diffraction (LXRD), SXRD boasts a significantly brighter beam—approximately 100 million times brighter than its laboratory-based counterpart—alongside tunable monochromatic wavelengths and enhanced collimation. Consequently, SXRD significantly enhances phase identification and quantitative analysis of minor materials [18,19]. Since the late 1900s, synchrotron techniques such as X-ray microdiffraction (μXRD), X-ray fluorescence (XRF), extended X-ray absorption fine structure (EXAFS), and transmission X-ray microscopy (TXM) have been utilized to examine manganese nodules from both soil and deep-sea environments. These techniques enable researchers to discern the spatial distribution of minerals at a microscale, explore microbiological contributions, and track compositional changes within heterogeneous (micro-) nodules during their formation [20,21,22,23,24,25].
Mn nodules can be classified into hydrogenetic- and diagenetic-origin, based on their metal precipitation processes, and exhibit variations in structure, chemical composition, and mineralogy during growth [5,26]. Generally, hydrogenetic nodules exhibit smooth surface textures, laminated structures, lower Mn/Fe ratios (<2.5), and higher contents of Co, Ce, Te, and Pt [5,27]. In contrast, diagenetic nodules exhibit rough surfaces, porous structures, higher Mn/Fe ratios (Mn/Fe ratio > 2.5), higher Li, Ni, Cu, and Ca contents, and lower Ce/Th and Ce/U ratios [4,12,26,28]. However, the mineral composition of Mn nodules has not been fully elucidated, although it is expected to vary depending on the growth process. An illustrative example lies in the Fe-Mn nodules atop the tabletops of the Magellan Seamounts, which are linked with Co-rich Fe-Mn crusts. These nodules exhibit distinct layers—an inner core surrounded by Fe- and Mn-rich layers—where the nuclei display varying Mn/Fe ratios, signaling potential diagenetic processes during the initial stages of nodule formation [7].
The mineral compositions of Mn nodules, particularly the abundances of the four manganese oxide minerals (todorokite, birnessite, buserite, and vernadite [17,29]), are difficult to determine due to their small particle sizes (0.01–0.1 μm) and low crystallinity, which prevents accurate analysis using conventional X-ray quantification methods. The chemical formulas of todorokite ((Na,Ca,K,Ba,Sr)1−x(Mn,Mg,Al)6O12·3–4H2O), birnessite ((Na,Ca,K)0.6(Mn4+,Mn3+)2O4·1.5H2O), vernadite ((Mn,Fe,Ca,Na)(O,OH)2·nH2O), and buserite (Na4Mn14O27·21H2O)) are similar. This similarity makes it challenging to differentiate them based on chemical composition, particularly through cation exchange within their structures. As a result, many previous studies on Mn nodules have provided only qualitative results for entire nodule samples or rough indications of the abundance of each mineral based on X-ray diffraction analysis graphs [27,30,31]. This study attempted to identify the minor mineral constituents in Mn nodules from the eastern and western Pacific Oceans and compare their chemical compositions. In cases where conventional quantitative analysis was challenging (i.e., for oxide minerals such as manganese oxides), we employed high-resolution data obtained through synchrotron X-ray diffraction analysis to perform semi-quantitative analysis. Through conducting this fundamental research, the advantages of high-resolution synchrotron X-ray diffraction analysis and additional applications of Mn mineral analysis are demonstrated.

2. Materials and Methods

2.1. Samples

A Manganese nodule (KC-8_EBS01_75, KC-8) used in this study was obtained using an epibenthic sledge from Korean Cluster-8 (KC-8, 149°34.786′ E, 17°06.125′ N, 1616 m below sea level, mbsl) of the Magellan Seamount in the Western Pacific Ocean during the R/V Isabu Expedition conducted by the Korea Institute of Ocean Science and Technology (KIOST) in 2021. Another manganese nodule (KODOS-10-2_BC_05, KODOS-10) was collected using a box core from the Clarion-Clipperton fracture zone (CCZ) of the eastern Pacific Ocean (131°39.796′ W, 10°09.340′ N, 5119 mbsl) during the R/V Onnuri Expedition in 2010 (Figure 1 and Figure 2).
A total of two Fe-Mn nodules with different shapes (KC-8: spherical and KODOS-10: semi-spherical) were selected for mineralogical and geochemical analyses (Figure 1). The collected samples were broken with a hammer and ground to a bulk powder using a ball mill (PM400, Retsch; 200 rpm, 60 min) to prepare a homogeneous sample. In total, we prepared five samples of each nodule for semi-quantification analysis.
Dehydrated samples were prepared by heating at ca. 90 °C for 12 h in air-dried condition to distinguish buserite and todorokite which have same basal reflection at 10 Å before dehydration of the buserite. We checked XRD patterns right after cooling down samples and checked again after 5 h. We found that the diffraction pattern was the same; therefore, we concluded the sample had not rehydrated during our measurement. Thermally stable buserite (buserite-II) as well as Mg-rich buserite has basal reflection of 9.6–9.7 Å. After examining our diffraction data, we could observe diffracted peak around not 9.6–9.7 Å but 10 Å. Also, basal peak was not broad, and no shoulder peak at 10 Å was detected. Therefore, we concluded that buserite-II might not exist, and the proportion of this mineral is negligible in our samples [32,33,34,35].

2.2. Chemical Analysis

The compacted bulk samples were placed on a carbon tape and coated with platinum for chemical analyses. The chemical compositions were observed via energy-dispersive X-ray spectroscopy (EDS) using a ZEISS Gemini 500 (ZEISS, Oberkochen, Baden-Württemberg, Germany) field-emission scanning electron microscope (FE-SEM) at Pusan National University. EDS analyses were performed at an acceleration voltage of 15 kV and an 8.5 mm working distance. The stoichiometry of elements in each sample was averaged by ten area scans (Table 1). The analysis results were plotted in a (Cu + Ni)-Fe-Mn ternary diagram, which was divided into hydrogenetic and diagenetic origins (Figure 3).

2.3. Synchrotron X-ray Powder Diffraction

High-resolution synchrotron X-ray powder diffraction data of the selected samples were measured at the 9 B beamline of the Pohang Accelerator Laboratory (PAL). The incident X-rays were vertically collimated using a mirror and monochromatized to wavelengths of 1.5419(1) or 1.5225(1) Å using a double-crystal Si(111) monochromator. The detector arm of the vertical scan diffractometer comprised six sets of Soller slits, flat Ge(111) crystal analyzers, anti-scatter baffles, and scintillation detectors, with each set separated by 20°. Each sample of ca. 0.2 g powder was prepared using the flat plate side loading method to avoid the preferred orientation, and the sample was then rotated approximately normal to the surface during the measurement to increase the sampling quality. Step scanning was performed at room temperature from 7° to 121° in two-theta with increments of 0.005° and 2° of overlap with the next detector bank. The samples were exposed for 4 s at each step to increase the degree of normalization of the diffracted beam intensity.

2.4. Phase Identification and Semi-Quantitative Analysis

A qualitative phase analysis of the minerals in the samples was performed using the Match!3 program (Crystal Impact, Bonn, Germany). The background curves were determined manually and graphically. For phase identification, the second-derivative method was used to search for Bragg’s peaks [36]. The “PDF-2” and “crystal open” mineral databases were used.
A semi-quantitative analysis of the selected manganate minerals was conducted by calculating the areas of the selected peaks [37,38]. The full width at half maximum (FWHM) of each peak was calculated using the profile fitting method to determine the integrated peak area for quantitative phase fraction analysis. The results of the qualitative and quantitative phase analyses are summarized in Table 2 and Table 3, respectively, with error values indicated in parentheses.

3. Results and Discussion

Table 1. Chemical compositions of bulk samples of Mn nodules. Errors are in parentheses.
Table 1. Chemical compositions of bulk samples of Mn nodules. Errors are in parentheses.
LocationElements (%)Mn/Fe
MnFeSiAlMgCaNaKClTiCoNiCu
KC-834.3 (1)26.0 (1)10.9 (1)4.4 (1)5.0 (1)4.2 (1)8.5
(1)
0.9
(1)
2.0
(1)
1.7
(1)
1.2
(1)
0.8
(1)
1.32
KODOS-1040.0 (1)6.4
(1)
17.8 (1)7.2 (1)6.1 (1)2.3 (1)10.2 (1)1.9
(1)
1.8
(1)
0.6
(1)
1.6
(1)
4.1
(1)
6.24
Southern
Pacific
Ocean a
16.116.17.33.01.32.01.80.9 1.00
Indian Ocean b24.47.110.02.91.71.91.10.2 3.42
East Siberian Sea c49.92.25.52.81.81.52.52.0 22.6
a Hein et al., 2013 [39], b Jauhari and Pattan, 2000, Pattan et al., 2001, Baturin and Dubinchuk, 2010, Pattan and Parthiban, 2011, Balaram et al., 2006 [40,41,42,43,44], c Yu et al., 2020 [38].
The chemical analysis results of two bulk powder samples, KC-8 and KODOS-10, are presented in Table 1. In the KC-8 sample, the most abundant element was Mn, accounting for 34.3% of the atomic fraction, followed by Fe (26.0%), Si (10.9%), Na (8.5%), Mg (5.0%), Al (4.4%), Ca (4.2%), and K (0.9%). In contrast, KODOS-10 showed Mn as the predominant element at 40.0%, followed by Si (17.8%), Na (10.2%), Al (7.2%), Fe (6.4%), Mg (6.1%), Ca (2.3%), and K (1.9%). The presence of Si and Al, albeit lower than Mn, suggests the coexistence of aluminosilicate minerals as minor phases. This implies the inflow of dissolved mineral sources or their presence on the sediment surface. The coexistence of these minerals with Mn and Fe may contribute to their potential dilution. Notably, the Fe content differed significantly between the two samples; KC-8 comprised 26.0% Fe, while KODOS-10 contained 6.4% Fe. Consequently, the Mn/Fe ratios were 1.32 for KC-8 and 6.24 for KODOS-10. These ratios are commonly utilized to differentiate between the two genetic formation origins of Mn nodules. Hydrogenetic and diagenetic formation is assumed when the ratio is below or above 5.0, respectively [45]. The Mn/Fe ratios observed in Mn nodules from the Southern Pacific Ocean, Indian Ocean, and East Siberian Sea were 1.00, 3.42, and 22.6, respectively (Table 1). These findings suggest that the nodules from the Southern Pacific Ocean and Indian Ocean were predominantly formed through hydrogenetic processes, while those from the East Siberian Sea originated primarily through diagenesis [2,38,39].
Figure 3 shows a (Cu + Ni)-Fe-Mn ternary diagram of Mn nodules from the Magellan Seamounts in the western Pacific Ocean [7]. The western Pacific Magellan Seamount cluster, formed around 120 million years ago during the Cretaceous period, and subsequently displaced northwest due to tectonic movements, offers ideal conditions for the development of Co-rich Fe-Mn crusts owing to its low sedimentation rate (0.4–4 mm/103 years) and obstruction of bottom currents by the seamounts [1]. In our previous results, we observed a transition from early diagenetic to hydrogenetic conditions, indicated by inner-to-outer part chemical variations (depicted as grey symbols). Additionally, the presence of todorokite primarily in the inner part, accompanied by high Mn/Fe ratios reaching up to 3.71, signifies the nodule’s exposure to diagenetic conditions during the initial stages of its formation. Subsequently, a shift in origin occurred at the tabletop of the seamount. Therefore, considering these findings, the KC-8 sample obtained from the same seamount is also deemed a mixed-type nodule, showcasing a similar evolutionary pattern in its formation as observed in previous analyses [7].
Figure 3. Ternary diagrams of Fe-Mn-(Cu + Ni) system to differentiate the genetic origins of the ferromanganese nodules. Grey and black colored symbols represent the results of our previous and present studies, respectively [7].
Figure 3. Ternary diagrams of Fe-Mn-(Cu + Ni) system to differentiate the genetic origins of the ferromanganese nodules. Grey and black colored symbols represent the results of our previous and present studies, respectively [7].
Minerals 13 01555 g003
Manganese nodules within the Clarion-Clipperton Zone (CCZ) in the Pacific Ocean are commonly classified as mixed-type nodules, yet their primary formations exhibit distinct differences. For instance, nodules sourced from Cook Island within the CCZ’s southeastern Pacific Ocean region are primarily formed through hydrogenetic precipitation. These nodules exhibit lower Mn/Fe ratios, approximately around one, signifying a hydrogenetic origin. In contrast, nodules from the northeastern CCZ region are characterized by a higher Mn/Fe ratio of approximately five, indicating that their composition is richer in diagenetic components. These differences underscore the diverse origins and predominant formations of manganese nodules within various regions of the CCZ [39,46].
The stacked synchrotron X-ray diffraction patterns are shown in Figure 4, and the phase identification results are summarized in Table 2. Diffraction peaks belonging to quartz (SiO2), buserite (Na4Mn14O27·21H2O), and vernadite ((Mn,Fe,Ca,Na)(O,OH)2·nH2O) were predominant, and feldspar such as anorthite (Ca[Si,Al]2O8) was present as minor a constituent in all samples. Todorokite ((Na,Ca,K,Ba,Sr)1−x(Mn,Mg,Al)6O12·3–4H2O), birnessite ((Na,Ca,K)0.6(Mn4+,Mn3+)2O4·1.5H2O) and illite (K0.65Al2.0(Al0.65Si3.35O10)(OH)2) were observed only in the KODOS-10 sample. According to Park et al., todorokite can be included in KC-8 sample as nuclei. However, we did not find todorokite in this work because the fraction of todorokite in the bulk powder sample was too low to be observed by X-ray diffraction [7]. The peak intensity of vernadite (observed at 36.5° in two-theta, coexisting with buserite) appears notably higher in KC-8 compared to KODOS-10. This disparity suggests that vernadite is predominantly formed in a hydrogenetic environment, primarily through the direct precipitation of Fe oxyhydroxide. This higher intensity in KC-8 implies a greater prevalence of hydrogenetic conditions favoring vernadite formation.
Figure 4. Stacked synchrotron X-ray powder diffraction patterns of the KC-8 and KODOS-10 samples.
Figure 4. Stacked synchrotron X-ray powder diffraction patterns of the KC-8 and KODOS-10 samples.
Minerals 13 01555 g004
Table 2. Phase identification results of KC-8 and KODOS-10.
Table 2. Phase identification results of KC-8 and KODOS-10.
KC-8KODOS-10
QuartzQuartz
FeldsparFeldspar
BuseriteTodorokite
VernaditeBuserite
Vernadite
Birnessite
Illite
In the case of KODOS-10, even after heating, a distinct peak at 9° persisted, identified as originating from todorokite. The notable presence of this peak aligns with the high Mn/Fe ratio (6.24), indicative of its formation primarily under diagenetic conditions. The composition of Mn nodules predominantly comprises todorokite, buserite, birnessite, and vernadite. These findings align consistently with earlier mineralogical studies conducted in the CCZ region, highlighting similar mineralogical compositions such as Fe-vernadite, todorokite, feroxyhyte, quartz, feldspar, and clay minerals [27]. They are mainly composed of manganate minerals such as 10 Å manganate minerals (e.g., 10 Å vernadite, todorokite, and buserite) and 7 Å manganate minerals (e.g., 7 Å vernadite and birnessite) [47,48].
The manganese layers in these minerals are formed by MnO6 octahedra interconnected through shared edges. Buserite, birnessite, and vernadite fall into the category of phyllomanganates, while todorokite belongs to tectomanganates. Buserite shares a similar framework with birnessite, allowing its Mn layer to readily expand or collapse based on the cation and hydration levels within its interlayers. Typically, the d-spacing of the Mn-layers (d(001)) in buserite measures around 10 Å, observed at a peak of 9° two-theta, and decreases to 7 Å at a peak of 12° two-theta when partially dehydrated to a hydration level akin to birnessite. The inflexible pore size of todorokite also measures around 10 Å, reflected in its (001) peak, similar in size and interlayer distance to buserite. Consequently, both todorokite and buserite exhibit their d(001) peak at the same two-theta (9°) in our samples. To achieve precise quantitative analysis and differentiate between todorokite and buserite, all samples underwent heating at 90 °C to partially dehydrate buserite and separate its (001) peak from the todorokite peak. For KC-8, the peak at 9° completely shifted to 12° after heating, indicating the exclusive presence of buserite in this sample. However, in KODOS-10, the intensity at 9° decreased while the peak at 12° increased after heating, indicating the coexistence of todorokite and buserite.
The semi-quantitative analysis results of the phase fractions of the selected manganate minerals (buserite, vernadite, and todorokite) based on comparing selected peak area are shown in Figure 5 and summarized in Table 3. The manganate in the KC-8 consisted mostly of vernadite, ranging from 94(5)%–100(4)%, and trace amounts of buserite, ranging from 6(1)%–0%. In contrast, while vernadite was still a major manganate in the KODOS-10, ranging from 47(1)%–56(2)%, the phase fractions of buserite and todorokite increased from 33.6(4)%–40.1(3)% and 10(3)%–16.3(8)%, respectively. In our previous study, Mn nodules in the tabletops of the OSM-9 Magellan seamount in the western Pacific Ocean included vernadite and todorokite. Diagenesis was initially dominant, while hydrogenesis gradually became more dominant over time. However, we only observed vernadite in our XRD results, and we consequently expect that the KC-8 did not undergo diagenetic processes or was formed after hydrogenetic conditions became dominant. In the KODOS-10, the vernadite and todorokite coexist. Like most of the nodules of this area, the nodules in this study were also mixed type nodules. The observed discrepancy in the mineral composition of the KODOS-10 sample could potentially be attributed to imperfect homogenization during sample preparation or indicate notable heterogeneity within the nodules. To delve deeper into comprehending the formation process specific to KODOS-10, our forthcoming study will systematically explore variations between its inner and outer parts. This investigation will employ a range of methodologies, including an age-dating method, chemical analysis, electron micro diffraction, and X-ray diffraction measurements. By scrutinizing these differences, we aim to gain a detailed and comprehensive understanding of the developmental mechanisms underlying the KODOS-10 nodules.
Figure 5. Ternary diagrams of phase fraction of buserite, todorokite, and vernadite resulting from semi-quantification analysis.
Figure 5. Ternary diagrams of phase fraction of buserite, todorokite, and vernadite resulting from semi-quantification analysis.
Minerals 13 01555 g005
Table 3. Phase fraction (%) of manganates in the KC-8 and the KODOS-10. Estimated standard deviation is in parentheses.
Table 3. Phase fraction (%) of manganates in the KC-8 and the KODOS-10. Estimated standard deviation is in parentheses.
Sample Sum
No.TodorokiteBuseriteVernadite
KC-8
(KC-8_1_EBS01_75)
101.0(1)99(4)100(2)
201(1)99(3)100(2)
300100(4)100(4)
406(1)94(5)100(3)
501(1)99(3)100(2)
KODOS-10
(KODOS10-2_BC-05)
116.3(8)36.7(9)47(2)100.0(8)
210(3)40(2)50(2)100(2)
310.5(2)33.6(4)56(2)100.1(9)
412.4(3)39.8(5)48(1)100.2(6)
512.6(2)40.1(3)47(1)99.7(5)
Figure 6 displays the powder diffraction patterns of KC-8 and KODOS-10 using Lab-XRD and synchrotron-XRD (LXRD and SXRD), respectively, aiming to highlight differences in peak width and signal-to-background ratios. In LXRD, the background signal appears notably higher compared to SXRD. This increased background signal is typically attributed to air scattering, fluorescence, and non-crystallized sample components. SXRD benefits from an exceptionally bright beam, approximately 100 million times stronger than laboratory-sourced X-rays. It employs a Soller slit and crystal analyzer before the diffractometer to filter out noise signals and elevate the sample-to-background ratio. When utilizing SXRD, we observe distinct peaks, notably around 12° and 18° of two-theta in the KC-8 sample, enabling precise identification and semi-quantification of minor phases within the samples. A comparison of Full Width at Half Maximum (FWHM) for selected peaks is summarized in Table 4, revealing changes in peak width based on two-theta degrees when using SXRD and LXRD. While the FWHM at 9° remains similar between both XRD types, ranging from 0.53(2) to 0.59(3), the FWHM at 36.5° significantly differs, spanning 0.99(3) to 3.49(12). Moreover, the FWHM values obtained from SXRD are consistently lower than those from LXRD. The phenomenon of overlapping Bragg peaks due to emission of characteristic X-rays Kα1 and Kα2 is observed in LXRD, especially impacting semi-crystallized minerals like vernadite. This overlap, especially at high angles, may lead to errors in semi-quantitative analysis based on peak area calculations. Conversely, SXRD employs monochromatic X-rays, resulting in sharper peak widths and enabling more accurate semi-quantitative analysis.
Table 4. Full width at half maximum (FWHM) of selected peaks.
Table 4. Full width at half maximum (FWHM) of selected peaks.
SampleMeasurementsFWHM (°)
9° at 2θ12°26.5°36.5°
KC-8Synchrotron-XRD0.56(5)1.75(17)0.064(4)0.99(3)
Lab-XRD0.59(3)-0.089(3)1.21(3)
KODOS-10Synchrotron-XRD0.53(2)0.33(3)-1.54(3)
Lab-XRD0.58(6)0.48(4)-3.49(12)

4. Conclusions

We conducted geochemical and mineralogical analyses on two bulk mixed-type Mn nodules sourced from the western and eastern Pacific Oceans. The objective was to discern differences in their genetic origins, mineral compositions, and the prevalence of various manganates. The Fe content varied between the samples, resulting in Mn/Fe ratios of 1.32 for KC-8 and 6.24 for KODOS-10. This suggests predominant influences of hydrogenetic and diagenetic conditions in the formation of KC-8 and KODOS-10, respectively. Utilizing high-resolution synchrotron X-ray diffraction (SXRD), both samples exhibited quartz, vernadite, buserite, and feldspar. However, KODOS-10 uniquely displayed birnessite and illite. Phase fractions revealed vernadite predominance above 94(5)% in KC-8, while KODOS-10 displayed a vernadite fraction ranging from 47(1)% to 56(2)%. Additionally, KODOS-10 exhibited higher proportions of buserite and todorokite, constituting 33.6(4)% to 40.1(3)%. The enhanced brightness and signal-to-background ratio of the SXRD facilitated clear observation of minor phase fractions, approximately 1%, a level not easily detected using traditional methods. Moreover, the Full Width at Half Maximum (FWHM) in SXRD was lower than that in LXRD due to differences in X-ray characteristics. We anticipate our findings will contribute fundamental insights into understanding the origin of Mn nodules in the Pacific Ocean and their distribution in seamounts. To delve deeper into nodule origins, we plan to conduct a detailed analysis of systematic differences between the inner and outer nodule sections of KODOS-10. This investigation will utilize an age-dating method, chemical analyses, electron micro diffraction, and X-ray diffraction measurements.

Author Contributions

Conceptualization, J.J. and Y.L.; methodology, S.L., J.K., K.H., Y.K., K.Y., H.A., J.P. and D.S.; software, S.L., H.A. and J.P.; validation, J.K., K.H., Y.K. and K.Y.; formal analysis, S.L., H.A. and J.P.; investigation, J.J., K.Y., D.S. and Y.L.; resources, J.J.; writing—original draft preparation, J.J. and Y.L.; writing—review and editing, J.J. and Y.L.; project administration, J.J. and Y.L.; funding acquisition, J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was part of the project titled ‘Deep sea and low-temperature environments, biogeochemical elemental cycle, microbe-mineral interaction, paleoenvironment, and Fe source’, funded by the Korea Institute of Ocean Science & Technology project (PE9996F). The APC was also funded by above project.

Data Availability Statement

Raw data are available from the corresponding authors upon request.

Acknowledgments

We are thankful for the extraordinary efforts of the scientists and crew members of the KIOST.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Glasby, G.P.; Ren, X.; Shi, X.; Pulyaeva, I.A. Co–rich Mn crusts from the Magellan Seamount cluster: The long journey through time. Geo-Mar. Lett. 2007, 27, 315–323. [Google Scholar] [CrossRef]
  2. Hein, J.R.; Koschinsky, A. 13.11–Deep-Ocean Ferromanganese Crusts and Nodules. In Treatise on Geochemistry, 2nd ed.; Holland, H.D., Turekian, K.K., Eds.; Elsevier: Oxford, UK, 2014; pp. 273–291. [Google Scholar] [CrossRef]
  3. Zhang, W.; Cheng, C.Y. Manganese metallurgy review. Part I: Leaching of ores/secondary materials and recovery of electrolytic/chemical manganese dioxide. Hydrometallurgy 2007, 89, 137–159. [Google Scholar] [CrossRef]
  4. Hein, J.; Koschinsky, A.; Kuhn, T. Deep-ocean polymetallic nodules as a resource for critical materials. Nat. Rev. Earth Environ. 2020, 1, 158–169. [Google Scholar] [CrossRef]
  5. Kuhn, T.; Wegorzewski, A.; Rühlemann, C.; Vink, A. Composition, Formation, and Occurrence of Polymetallic Nodules. In Deep-Sea Mining; Springer: Cham, Switzerland, 2017; pp. 23–63. [Google Scholar] [CrossRef]
  6. Murton, B.; Parson, L.; Hunter, P.; Miles, P. Global Non-Living Resources on the Extended Continental Shelf: Prospects at the Year 2000; International Seabed Authority: Kingston, Jamaica, 2001. [Google Scholar]
  7. Park, J.; Jung, J.; Ko, Y.; Lee, Y.; Yang, K. Reconstruction of the Paleo-Ocean Environment Using Mineralogical and Geochemical Analyses of Mixed-Type Ferromanganese Nodules from the Tabletop of Western Pacific Magellan Seamount. Geochem. Geophys. Geosystems 2023, 24, e2022GC010768. [Google Scholar] [CrossRef]
  8. 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]
  9. Burns, R.G.; Mee Burns, V. Mechanism for nucleation and growth of manganese nodules. Nature 1975, 255, 130–131. [Google Scholar] [CrossRef]
  10. Burns, V.; Burns, R. Observations of processes leading to the uptake of transition metals in manganese nodules. Genèse Nodules Manganèse 1979, 289, 387–404. [Google Scholar]
  11. Burns, R.; Burns, V. Mineralogy. In Marine Manganese Deposits: Elsevier Oceanography Series; Elsevier: Amsterdam, The Netherlands, 1977; Volume 15, pp. 185–248. [Google Scholar]
  12. Manceau, A.; Lanson, M.; Takahashi, Y. Mineralogy and crystal chemistry of Mn, Fe, Co, Ni, and Cu in a deep-sea Pacific polymetallic nodule. Am. Mineral. 2014, 99, 2068–2083. [Google Scholar] [CrossRef]
  13. Wegorzewski, A.V.; Grangeon, S.; Webb, S.M.; Heller, C.; Kuhn, T. Mineralogical transformations in polymetallic nodules and the change of Ni, Cu and Co crystal-chemistry upon burial in sediments. Geochim. Cosmochim. Acta 2020, 282, 19–37. [Google Scholar] [CrossRef]
  14. Atkins, A.L.; Shaw, S.; Peacock, C.L. Release of Ni from birnessite during transformation of birnessite to todorokite: Implications for Ni cycling in marine sediments. Geochim. Cosmochim. Acta 2016, 189, 158–183. [Google Scholar] [CrossRef]
  15. Muñoz-Royo, C.; Peacock, T.; Alford, M.H.; Smith, J.A.; Le Boyer, A.; Kulkarni, C.S.; Lermusiaux, P.F.J.; Haley, P.J.; Mirabito, C.; Wang, D.; et al. Extent of impact of deep-sea nodule mining midwater plumes is influenced by sediment loading, turbulence and thresholds. Commun. Earth Environ. 2021, 2, 148. [Google Scholar] [CrossRef]
  16. González, F.; Somoza, L.; Hein, J.; Medialdea, T.; León, R.; Urgorri, V.; Reyes, J.; Martín-Rubí, J. Phosphorites, Co-rich Mn nodules, and Fe-Mn crusts from Galicia Bank, NE Atlantic: Reflections of Cenozoic tectonics and paleoceanography. Geochem. Geophys. Geosystems 2016, 17, 346–374. [Google Scholar] [CrossRef]
  17. Vereshchagin, O.; Perova, E.; Brusnitsyn, A.; Ershova, V.; Khudoley, A.; Shilovskikh, V.; Molchanova, E. Ferro-manganese nodules from the Kara Sea: Mineralogy, geochemistry and genesis. Ore Geol. Rev. 2019, 106, 192–204. [Google Scholar] [CrossRef]
  18. Kamitsubo, H. SPring-8 program. J. Synchrotron Radiat. 1998, 5, 162–167. [Google Scholar] [CrossRef] [PubMed]
  19. Withers, P.J. Synchrotron X-ray Diffraction. In Practical Residual Stress Measurement Methods; Wiley: Hoboken, NJ, USA, 2013; pp. 163–194. [Google Scholar] [CrossRef]
  20. Brown, G.E.; Foster, A.L.; Ostergren, J.D. Mineral surfaces and bioavailability of heavy metals: A molecular-scale perspective. Proc. Natl. Acad. Sci. USA 1999, 96, 3388–3395. [Google Scholar] [CrossRef] [PubMed]
  21. Manceau, A.; Marcus, M.A.; Tamura, N. Quantitative speciation of heavy metals in soils and sediments by synchrotron X-ray techniques. Rev. Mineral. Geochem. 2002, 49, 341–428. [Google Scholar] [CrossRef]
  22. Manceau, A.; Tamura, N.; Celestre, R.S.; MacDowell, A.A.; Geoffroy, N.; Sposito, G.; Padmore, H.A. Molecular-scale speciation of Zn and Ni in soil ferromanganese nodules from loess soils of the Mississippi Basin. Environ. Sci. Technol. 2003, 37, 75–80. [Google Scholar] [CrossRef]
  23. Yu, X.-L.; Fu, Y.-N.; Brookes, P.C.; Lu, S.-G. Insights into the formation process and environmental fingerprints of iron–manganese nodules in subtropical soils of China. Soil Sci. Soc. Am. J. 2015, 79, 1101–1114. [Google Scholar] [CrossRef]
  24. Jiang, X.D.; Zhao, X.; Chou, Y.M.; Liu, Q.S.; Roberts, A.P.; Ren, J.B.; Sun, X.M.; Li, J.H.; Tang, X.; Zhao, X.Y.; et al. Characterization and Quantification of Magnetofossils Within Abyssal Manganese Nodules from the Western Pacific Ocean and Implications for Nodule Formation. Geochem. Geophys. Geosystems 2020, 21, e2019GC008811. [Google Scholar] [CrossRef]
  25. Vodyanitskii, Y.N. The composition of Fe-Mn nodules as determined by synchrotron X-ray analysis (Review of publications). Eurasian Soil Sci. 2006, 39, 147–156. [Google Scholar] [CrossRef]
  26. Bau, M.; Koschinsky, A.; Dulski, P.; Hein, J.R. Comparison of the partitioning behaviours of yttrium, rare earth elements, and titanium between hydrogenetic marine ferromanganese crusts and seawater. Geochim. Cosmochim. Acta 1996, 60, 1709–1725. [Google Scholar] [CrossRef]
  27. Wegorzewski, A.; 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]
  28. Halbach, P. Processes controlling the heavy metal distribution in Pacific ferromanganese nodules and crusts. Geol. Rundsch. 1986, 75, 235–247. [Google Scholar] [CrossRef]
  29. Glasby, G. The mineralogy of manganese nodules from a range of marine environments. Mar. Geol. 1972, 13, 57–72. [Google Scholar] [CrossRef]
  30. Reykhard, L.Y.; Shulga, N. Fe-Mn nodule morphotypes from the NE Clarion-Clipperton Fracture Zone, Pacific Ocean: Comparison of mineralogy, geochemistry and genesis. Ore Geol. Rev. 2019, 110, 102933. [Google Scholar] [CrossRef]
  31. Lee, S.; Xu, H. XRD and TEM studies on nanophase manganese oxides in freshwater ferromanganese nodules from Green Bay, Lake Michigan. Clays Clay Miner. 2016, 64, 523–536. [Google Scholar] [CrossRef]
  32. Bish, D.L.; Post, J.E. Thermal behavior of complex, tunnel-structure manganese oxides. Am. Mineral. 1989, 74, 177–186. [Google Scholar]
  33. Shen, Y.F.; Zerger, R.P.; DeGuzman, R.N.; Suib, S.L.; McCurdy, L.; Potter, D.I.; O’Young, C.L. Manganese Oxide Octahedral Molecular Sieves: Preparation, Characterization, and Applications. Science 1993, 260, 511–515. [Google Scholar] [CrossRef]
  34. Ching, S.; Krukowska, K.S.; Suib, S.L. A new synthetic route to todorokite-type manganese oxides. Inorganica Chim. Acta 1999, 294, 123–132. [Google Scholar] [CrossRef]
  35. Uspenskaya, T.Y.; Gorshkov, A.; Sivtsov, A. Mineralogy and internal structure of Fe-Mn nodules from the Clarion-Clipperton fracture zone. Int. Geol. Rev. 1987, 29, 363–371. [Google Scholar] [CrossRef]
  36. Schreiner, W.N.; Jenkins, R. A Second Derivative Algorithm for Identification of Peaks in Powder Diffraction Patterns. Adv. X-ray Anal. 1979, 23, 287–293. [Google Scholar] [CrossRef]
  37. Biscaye, P.E. Mineralogy and Sedimentation of Recent Deep-Sea Clay in the Atlantic Ocean and Adjacent Seas and Oceans. GSA Bull. 1965, 76, 803–832. [Google Scholar] [CrossRef]
  38. Yu, H.J.; Shin, E.J.; Koo, H.J.; Cho, H.G. Semi-quantitative Analysis of Manganese Oxide Mineral in Manganese Nodule from the East Siberian Sea. Korean J. Mineral. Petrol. 2020, 33, 427–437. [Google Scholar]
  39. 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. Ore Geol. Rev. 2013, 51, 1–14. [Google Scholar] [CrossRef]
  40. Jauhari, P.; Pattan, J. Ferromanganese nodules from the central Indian Ocean basin. In Handbook of Marine Mineral Deposits; Routledge: London, UK, 2017; pp. 171–196. [Google Scholar]
  41. Pattan, J.; Rao, C.M.; Migdisov, A.; Colley, S.; Higgs, N.; Demidenko, L. Ferromanganese nodules and their associated sediments from the Central Indian Ocean Basin: Rare earth element geochemistry. Mar. Georesources Geotechnol. 2001, 19, 155–165. [Google Scholar] [CrossRef]
  42. Baturin, G.; Dubinchuk, V. On the composition of ferromanganese nodules of the Indian Ocean. Proc. Dokl. Earth Sci. 2010, 343, 1179–1184. [Google Scholar] [CrossRef]
  43. Pattan, J.N.; Parthiban, G. Geochemistry of ferromanganese nodule–sediment pairs from Central Indian Ocean Basin. J. Asian Earth Sci. 2011, 40, 569–580. [Google Scholar] [CrossRef]
  44. Balaram, V.; Mathur, R.; Banakar, V.; Hein, J.R.; Rao, C.; Rao, T.G.; Dasaram, B. Determination of the platinum–group elements (PGE) and gold (Au) in manganese nodule reference samples by nickel sulfide fire-assay and Te coprecipitation with ICP-MS. Indian J. Geo-Mar. Sci. 2006, 35, 7–16. [Google Scholar]
  45. 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]
  46. Hein, J.R.; Spinardi, F.; Okamoto, N.; Mizell, K.; Thorburn, D.; Tawake, A. Critical metals in manganese nodules from the Cook Islands EEZ, abundances and distributions. Ore Geol. Rev. 2015, 68, 97–116. [Google Scholar] [CrossRef]
  47. Wegorzewski, A.V.; Kuhn, T.; Dohrmann, R.; Wirth, R.; Grangeon, S. Mineralogical characterization of individual growth structures of Mn-nodules with different Ni+Cu content from the central Pacific Ocean. Am. Mineral. 2015, 100, 2497–2508. [Google Scholar] [CrossRef]
  48. Su, R.; Sun, F.; Li, X.; Chu, F.; Sun, G.; Li, J.; Wang, H.; Li, Z.; Zhang, C.; Zhang, W.; et al. Diverse early diagenetic processes of ferromanganese nodules from the eastern Pacific Ocean: Evidence from mineralogy and in-situ geochemistry. Int. Geol. Rev. 2022, 65, 2131–2147. [Google Scholar] [CrossRef]
Figure 1. Sampling location of Mn nodules from the Western (KC-8) and Eastern (KODOS-10) Pacific Oceans.
Figure 1. Sampling location of Mn nodules from the Western (KC-8) and Eastern (KODOS-10) Pacific Oceans.
Minerals 13 01555 g001
Figure 2. Deep Sea Camera (DSC) image of ferromanganese nodules (black dots) on tabletop of the KC-8 seamount in the Magellan Seamount Cluster.
Figure 2. Deep Sea Camera (DSC) image of ferromanganese nodules (black dots) on tabletop of the KC-8 seamount in the Magellan Seamount Cluster.
Minerals 13 01555 g002
Figure 6. Stacked plots of powder diffraction patterns using laboratory- and synchrotron-based XRD.
Figure 6. Stacked plots of powder diffraction patterns using laboratory- and synchrotron-based XRD.
Minerals 13 01555 g006
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

Jung, J.; Lee, S.; Yang, K.; Kim, J.; Hyeong, K.; Ko, Y.; An, H.; Park, J.; Seoung, D.; Lee, Y. Semi-Quantitative Analyses of Ferromanganese Nodules from the Pacific Ocean Using Synchrotron X-ray Powder Diffraction. Minerals 2023, 13, 1555. https://doi.org/10.3390/min13121555

AMA Style

Jung J, Lee S, Yang K, Kim J, Hyeong K, Ko Y, An H, Park J, Seoung D, Lee Y. Semi-Quantitative Analyses of Ferromanganese Nodules from the Pacific Ocean Using Synchrotron X-ray Powder Diffraction. Minerals. 2023; 13(12):1555. https://doi.org/10.3390/min13121555

Chicago/Turabian Style

Jung, Jaewoo, Soojin Lee, Kiho Yang, Jonguk Kim, Kiseong Hyeong, Youngtak Ko, Hyeonho An, Jinsub Park, Donghoon Seoung, and Yongmoon Lee. 2023. "Semi-Quantitative Analyses of Ferromanganese Nodules from the Pacific Ocean Using Synchrotron X-ray Powder Diffraction" Minerals 13, no. 12: 1555. https://doi.org/10.3390/min13121555

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