Next Article in Journal
The Flotation of Kyanite and Sillimanite with Sodium Oleate as the Collector
Previous Article in Journal
The Influence of Impurity Monovalent Cations Adsorption on Reconstructed Chalcopyrite (001)-S Surface in Leaching Process
Previous Article in Special Issue
Composition and Formation of Gabbro-Peridotite Hosted Seafloor Massive Sulfide Deposits from the Ashadze-1 Hydrothermal Field, Mid-Atlantic Ridge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Marine Phosphorites as Potential Resources for Heavy Rare Earth Elements and Yttrium

1
Pacific Coastal and Marine Science Center, U.S. Geological Survey, 2885 Mission St., Santa Cruz, CA 95060, USA
2
Department of Physics and Earth Sciences, Jacobs University, Bremen, P.O. Box 750561, Bremen D-28725, Germany
3
Ocean Sciences, University of California at Santa Cruz, Santa Cruz, CA 95064, USA
4
Department of Geology and Geophysics, School of Ocean and Earth Science and Technology, University of Hawaii, 1680 East West Rd., POST 701, Honolulu, HI 96822, USA
5
Chief Operations Officer, Chatham Rock Phosphate, 93 The Terrace, Wellington 6011, New Zealand
*
Author to whom correspondence should be addressed.
Minerals 2016, 6(3), 88; https://doi.org/10.3390/min6030088
Submission received: 13 June 2016 / Revised: 10 August 2016 / Accepted: 17 August 2016 / Published: 29 August 2016
(This article belongs to the Special Issue Marine Minerals: From Genesis to Resources)

Abstract

:
Marine phosphorites are known to concentrate rare earth elements and yttrium (REY) during early diagenetic formation. Much of the REY data available are decades old and incomplete, and there has not been a systematic study of REY distributions in marine phosphorite deposits that formed over a range of oceanic environments. Consequently, we initiated this study to determine if marine phosphorite deposits found in the global ocean host REY concentrations of high enough grade to be of economic interest. This paper addresses continental-margin (CM) and open-ocean seamount phosphorites. All 75 samples analyzed are composed predominantly of carbonate fluorapatite and minor detrital and authigenic minerals. CM phosphorites have low total REY contents (mean 161 ppm) and high heavy REY (HREY) complements (mean 49%), while seamount phosphorites have 4–6 times higher individual REY contents (except for Ce, which is subequal; mean ΣREY 727 ppm), and very high HREY complements (mean 60%). The predominant causes of higher concentrations and larger HREY complements in seamount phosphorites compared to CM phosphorites are age, changes in seawater REY concentrations over time, water depth of formation, changes in pH and complexing ligands, and differences in organic carbon content in the depositional environments. Potential ore deposits with high HREY complements, like the marine phosphorites analyzed here, could help supply the HREY needed for high-tech and green-tech applications without creating an oversupply of the LREY.

Graphical Abstract

1. Introduction

Rare earth elements (REE) plus yttrium (REY) are known to concentrate in marine phosphorite deposits during their formation, primarily through early diagenetic processes where carbonates are replaced by carbonate fluorapatite (CFA), which also precipitates in void space (e.g., [1,2]). However, much of the REY data available in the literature are decades old and incomplete, and there has not been a systematic study of REY distributions in marine phosphorite deposits that formed over a range of oceanic environments (e.g., [1,2,3]). Consequently, this study was initiated to determine the range of REY concentrations hosted by marine phosphorite deposits from the modern global ocean and whether the deposits are of high enough grades to be of economic interest.
Phosphorites in the global ocean occur in three general environments, continental margins (CM: shelf, slope, banks, and plateaus), seamounts, especially the old (Cretaceous) seamounts in the NW Pacific, and lagoon/insular deposits (e.g., [4,5]). Research presented here addresses only the CM and seamount phosphorites. Our global dataset includes typical shallow-water continental-margin, upwelling-zone phosphorites, and deep-water seamount phosphorites collected from the Pacific and Atlantic Oceans (Table S1, Figure 1).
Similar studies have been completed for land-based phosphorite deposits, with positive results as to the resource potential of REY as a byproduct or co-product of the focus phosphate mining (e.g., [6,7]). However, recovery of these land-based REY would require the addition of costly infrastructure and changes in extractive processing to the existing phosphate mining operations. For example, it would be beneficial to change the dissolution acid from sulfuric acid, which produces phosphogypsum as a waste product, to hydrochloric acid. This would alleviate processing a second solid for recovery of the REY partitioned into the phosphogypsum phase (e.g., [6,8]). Production of REY as a byproduct or co-product of phosphate mining needs to be considered in the early stages of planning for a new terrestrial or marine mining operation. With this in mind, three CM areas have been leased in the global ocean for exploration for marine phosphorites: Chatham Rise off New Zealand by Chatham Rock Phosphate Ltd. (Wellington, New Zealand); offshore Baja California, Mexico by Odyssey Marine Exploration and Exploraciones Oceanicas; and on the shelf off Namibia by Namibian Marine Phosphate Ltd. Again, consideration of REY as a byproduct at an early planning stage for a marine phosphorite operation would be warranted should extraction of the REY be determined to be economic.
The large land-based carbonatite REY ores contain more than 99% light REY (LREY) and new mines put into operation for recovery of the undersupplied heavy (HREY) as a consequence could create an oversupply of LREY to the market (e.g., [6]). Consequently, a search for deposits with the largest HREY/LREY ratios would help alleviate the supply problem and marine phosphorites might be able to fill that need. Here, we define the LREE as La through Sm and the HREY as Eu through Lu including Y. We use this division because there is a clear break in the abundance of the REY between Sm and Eu in typical large land-based REY deposits that supply most of the REY used worldwide. The average LREE complement from the carbonatite ores supply a resource about three orders of magnitude larger than the HREY complement (e.g., [9]).

2. Samples and Methods

2.1. Sample Collection

Phosphorites were collected from dredge hauls taken during various cruises, except Johnson Sealink (JSL) samples were collected by manned submersible. Phosphorite samples were selected based on geographic location and major element concentrations. Our global dataset includes typical shallow-water (<1000 m) CM, upwelling-zone phosphorites from the east Pacific (southern California Borderland, n = 10; Peru margin, n = 10), Chatham Rise in the southwest Pacific (n = 15), Blake Plateau in the NW Atlantic (n = 10), and a NE Atlantic bank (n = 1 leached sample), and typical deep-water (1400 to ~3000 m) seamount phosphorites collected throughout the global ocean (n = 12 untreated and n = 18 leached samples to remove calcite; Table S1, Figure 1 and Figure 2).
Several seamount samples (Table S1) from water depths >3000 m may have been displaced downslope and collected as talus. In situ phosphatization on seamounts usually occurred at water depths of less than 2500 m, rarely to 3000 m [4,10].
The southern California borderland CM phosphorite samples are composed of breccia and sandstone cemented and replaced by CFA (Figure S1A). These Borderland deposits include phosphorite slabs and phosphorite fragments that vary from pale-brown to dark brown and black, and the surface is generally polished and smooth. The Peru CM phosphorite samples are cobble conglomeritic hardground crusts that form slabs and fragments derived from cementation and replacement of carbonates by CFA, and that vary from black to pale brown, dark brown, and green-brown. A rough surface texture was inherited from the host clastic rock, and a few samples are polished. Several samples contain many small vugs, some filled with glauconite. These and similar samples from the region are described by [11,12]. The Chatham Rise phosphorite samples are polished pebbles of carbonates that were cemented and replaced by CFA (Figure S1B). The deposit was initially composed of phosphorite slabs that were subsequently fragmented by iceberg gouging during sea-level low-stands [13]. The phosphorite pebbles are typically dark brown and interspersed with off-white shell fragments. Blake Plateau CM phosphorite samples are composed of carbonates and breccia cemented and replaced by CFA, with predominantly well-polished phosphorite fragments and some phosphorite slabs. Samples are pale brown, pinkish-brown, and dark brown, and surfaces vary from polished and smooth to flaky. The seamount phosphorite samples are predominantly pelagic and reef carbonates cemented and replaced by CFA. These deposits are massive to layered fine-grained replaced foraminiferal sands, or replaced breccias (replaced reef material), and vary from pale brown, dark brown, pinkish pale brown, to gray, and typically have a Fe-Mn crust on the upper surface (Figure 2). Some samples are porous and others are very dense.

2.2. Sample Preparation and Analyses

One CM and 18 seamount samples were leached using 0.5 M triammonium citrate and 30% H2O2 to remove calcite and organic matter [4], leaving pure CFA in most of those samples based on X-ray diffraction (XRD) analysis (Table 1 and Table 2); trace minerals below detection limit could be present. Those samples were leached so that the pure CFA could be dated using Sr isotopes [4]. In addition, some samples from all the locations were partially impregnated or coated with impurities such as Fe and Mn oxides, which were mechanically removed prior to mineralogical and chemical analyses. The phosphorite samples were then ground to <75 μm prior to XRD and chemical analyses, using both a shatter box, where the sample was contained in a ceramic dish, and an IKA A11 analytical mill.
X-ray diffraction mineralogy on all samples was completed using a Philips diffractometer (Philips, Amsterdam, The Netherlands) with Cu Kα radiation and graphite monochromator. Semi-quantitative mineral percentages were calculated by using peak intensities and weighting factors relative to quartz set as 1 [14,15]. The detection limit for each mineral falls between about 0.2% and 1.0%.
Chemical analyses of major and minor elements for all samples were done at SGS analytical laboratory, Canada. The concentrations of the 10 major elements were determined by XRF on a borate-fused disk; Ba, Cr, Cu, Li, Ni, S, Sr, V, Zn, and Zr by ICP using four-acid (HCl, HF, HNO3, and HClO4) digest, with resulting solutions dried and then dissolved in 1 mL of aqua regia and diluted to 10.0 g with 1% HNO3; Ag, As, Be, Bi, Cd, Cs, Ga, Hf, In, Mo, Pb, Rb, Sb, Sc, Se, Sn, Ta, Te, Tl, and W by ICP-mass spectrometry (ICP-MS) after the same four-acid digest procedure; Co, Nb, Th, U, V, Zr, and REY by ICP-MS using a Li-metaborate fused disk that was dissolved in weak HNO3; CO2 by combustion and infrared spectroscopy; and F and Cl by specific-ion electrode. For leached phosphorite samples only, CO2 concentrations were determined by coulometric titration, and S by combustion and infrared spectroscopy. The Pearson correlation coefficient was used to calculate coefficient matrices for the chemical data.
REY plots were normalized to shale (PAAS, Post-Archean Australian Shale [16]). The Ce anomaly was calculated as Ce * = CeN/PrN with both PAAS- and chondrite-normalized [17] values; Ce * = 2CeN/LaN + PrN was not used because of the strong positive La anomaly (see [18]).

3. Results

This study provides data for high-grade phosphorite deposits from the various global-ocean settings described in the Samples and Methods section. Phosphatization forms a continuum from pervasive, which produces nearly pure CFA rocks, to rocks showing only minor phosphatization and detrital minerals or biogenic carbonates dominate (Figure 3). Many bulk untreated phosphorite samples contain only CFA, as do the leached samples (Table 1 and Table 2).

3.1. Mineralogy

The seamount phosphorites are composed predominantly of CFA, with lesser amounts of biogenic calcite, and detrital quartz and feldspars (Table 1); other minor to trace minerals are also present. The seamount samples selected for leaching were composed predominantly of CFA and calcite. Once leached, 13 of the 17 contained only CFA, two also contained phillipsite, and two have minor quartz (Table 1). The mineralogy of the CM phosphorites varies with location. The California borderland phosphorites are the most pure, with minor detrital and authigenic mineral fractions in 60% of the samples and moderate amounts of detrital minerals in 40% of the samples (Table 2). Chatham Rise and Black Plateau samples are the most carbonate-rich phosphorites, with calcite being a major phase in several samples. Detrital minerals and authigenic glauconite are also common in phosphorites from those two areas. The Peru margin phosphorites contain the highest detrital and authigenic mineral fractions (Table 2).

3.2. Geochemistry

The chemical composition supports the XRD mineralogy and shows that all samples are composed predominantly of CFA, based on P contents. The Ca content is distributed among the CFA, biogenic carbonate, and detrital minerals. The Ca/P ratios indicate that more calcite occurs in Chatham Rise and Blake Plateau phosphorites (3.7 and 3.6 respectively; Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8, Table S1; Figure 3) compared to the other CM and seamount phosphorites (2.6 to 2.7), especially the Blake Plateau samples, which also have the lowest mean Ca/CO2 ratio (2.45). Si and Al contents increase with increasing contents of biogenic silica, authigenic quartz, and detrital quartz and aluminosilicate minerals. Peru margin phosphorites have the highest aluminosilicate fraction and the seamount phosphorites the lowest, which is consistent with distance from detrital sources. Seamount phosphorites have Si/Al ratios (mean 3.0; 2.9 leached) similar to mean MORB and mean ocean crust (2.9–3.1), whereas all the CM deposits have Si/Al ratios (3.9–4.2) similar to upper continental crust (3.8–4.2), except Chatham Rise phosphorites (7.4). The Chatham Rise phosphorites show excess silica over typical continental detrital ratios, which reflect either biogenic silica (diatoms) or abundant quartz, the former being most likely. In general, Peru margin, Chatham Rise, and Blake Plateau phosphorite samples contain the highest non-phosphate components, predominantly detrital aluminosilicates, biogenic silica, and biogenic calcite respectively.

3.3. Rare Earth Elements (REY)

The continental-margin deposits have relatively low grades, with average total REY of 161 ppm (68–353 ppm, maximum 663 ppm; Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8, Table S2). Grades for the seamount phosphorite deposits are higher, averaging 727 ppm total REY (maximum 1992 ppm; Table 7 and Table 8, Table S3). However, the most interesting characteristic of both groups of phosphorites is the high percentages of the HREY. For the CM deposits, the %HREY ranges from 29% to 61% and averages 49%. For the seamount group, the average %HREY is 60% (range 47%–74%). The highest values come from the equatorial west Pacific seamounts (60%–74% HREY complement).
The Pacific and northeast Atlantic CM REY datasets are similar and differ significantly from the Blake Plateau dataset (Table 3, Table 4, Table 5 and Table 6). The Blake samples have the highest Mn concentration, which increases the concentrations of the Mn-hosted metals, such as Co, Ni, V, Pb, Zn, Zr, Cu, Mo, and REY. Consequently, Blake Plateau phosphorites have the highest REY contents compared to the other CM deposits, mean 353 ppm.
The highest concentrations of individual REY in all deposits are Y, La, and Nd in decreasing order, with Y significantly enriched over the others (Figure 4). The seamount phosphorites are enriched in all the REY except Ce compared to the CM phosphorites. The Blake Plateau and NE Atlantic datasets have the highest Ce contents. Besides Y, other HREY are significantly enriched in the seamount samples, Gd, Dy, Er, and Yb (Figure 4). Plots of PAAS-normalized REY show HREY enrichments relative to LREY, negative Ce anomalies (Ce *), and positive Y, Gd, and La anomalies (Figure 5). These characteristics generally reflect those of seawater, but the Ce * varies widely. Ce * increases in magnitude with increasing total REY contents, so the Peru margin and seamount phosphorites have the smallest and largest magnitude anomalies, respectively, compared to the other datasets.

3.4. Correlations

The presence of calcite and Mn oxides in the CM phosphorites changes the correlation coefficients for the CFA-hosted elements. The California and Peru margin samples have the lowest calcite and/or Fe-Mn oxides and show positive correlations among Ca, P, Sr, S, CO2, and %HREY, and at a lower statistical significance (95% confidence level: CL) with F, Cl, Mg, Sb, Cd, Mo, Se, U, Tl, and Ag. These elements can be considered to be associated for the most part with the CFA. The REY are fractionated among the CFA, aluminosilicate, and Mn-oxide minerals. With higher calcite contents, Ca no longer correlates with the other CFA elements (except CO2) at the 95% CL and has either a negative correlation with total REY (Chatham samples) or no correlation with total REY (Blake samples). ΣREY correlates at the 95%–99% CL predominantly with Mn and Fe-Mn-oxide associated elements in Blake samples. Mn in Blake samples correlates with Co, Ni, ΣREY, each REY (except Ce), Cu, Ti, Mo, Tl, W, Cd, Li, and Sc; coefficients are higher for the HREY than for the LREY.
The seamount samples show positive correlations among Ca, P, S, CO2, Na, F, Sr, U, and %HREY. These elements are considered to be associated with the CFA and as that phase increases so does the content of HREY. The REY are distributed among the CFA, Mn oxides, and aluminosilicate minerals. The elements that typically correlate with Si and Al in all the datasets are Fe, Na, K, Rb, Mg, Ti, Zr, Be, Cd, Th, Nb, and the LREY, which make up the silicate fraction of the phosphorites.

4. Discussion

4.1. Comparison of CM and Seamount Mineralization, and REY Mass Balance

It is well known that apatite is a key mineral in the acquisition of REY from a wide variety of fluids occurring in many different geological settings (e.g., [20,21]). Our data show that CM marine phosphorites have relatively low total REY concentrations with high complements of HREY. In comparison, seamount phosphorites have about four to six times higher individual REY concentrations than CM phosphorites, except for Ce, which is comparable, and very high HREY complements. The seamount deposits generally have lower detrital (mean 9.4% versus 12.1% for CM), biosilica, and biocalcite contents and consequently higher Ca and P contents, forming more pure CFA (mean 72% CFA) seamount phosphorite deposits compared to CM deposits (mean 54% CFA), based on formulas to calculate normative components from [22]. However, the higher Ca and P contents are only part of the story and do not fully explain the higher REY concentrations and higher complement of HREY in the seamount deposits. Based on mass-balance calculations, a small part of the REY content (8% for CM, 2% for seamount deposits) occurs in the detrital/authigenic mineral fraction and the remainder is associated with the CFA, and also minor concentrations with the Fe-Mn oxides in Blake Plateau samples and with calcite in samples from all areas. This is further reflected in the seamount phosphorite samples, which have the lowest detrital and authigenic fractions and the highest REY contents.

4.2. CFA Structure and Substitutions by REY

The REY typically exchange for Ca in the CFA structure; the LREY favor the Ca2 structural site and the HREY the Ca1 structural site, with the REY near Eu in ionic radius having no preference (e.g., [21]). Charge balance has to be maintained for substitutions in the CFA structure as there are many elements that can substitute for the Ca2+ and PO43−. Substitution of the trivalent REY for divalent Ca produces a charge imbalance that can be offset for example by the substitution of monovalent Na for Ca, silicate for phosphate, or charge-balancing anions such as F and Cl.
For the non-calcite bearing samples, the Ca/P ratios and CO2 contents are about the same for seamount and CM deposits, but Ca/CO2, P/CO2, Ca/Na, P/Na, and the magnitude of the negative Ce * are much lower for the CM phosphorites. These and other ratios indicate that there is comparatively less CO2, SO3, and Na relative to the main structural elements in the CFA of seamount phosphorites. The seamount CFA has more F and less Cl and Sr than the mean for the CFA fraction of the CM phosphorites. Of these relationships, less CO2 (CO3) and SO3 substitution for PO4, and higher F would aid in the charge balance created by the substitution of trivalent elements for the divalent Ca in the seamount CFA relative to the CM CFA. Although they can explain some variance, these chemical differences also cannot fully explain the different REY contents and LREY/HREY fractionations in CM versus seamount phosphorites.

4.3. Controls on REY Composition

There are four fundamental differences between the CM and seamount phosphorites. The seamount phosphorites are older and have therefore had more time for aging that might have allowed for greater REY adsorption, substitution, and replacement in the CFA. A second difference is the generally greater water depth of formation of the seamount phosphorites (1000 to ~3000 m) compared to CM deposits (10s to 100s m). The Pacific Cretaceous seamounts had subsided to near their present depths by the time the Paleogene and Miocene phosphorites had formed [4]. Dissolved REY, except Ce, increase with increasing water depth in the open Pacific Ocean (e.g., [23,24]), which would offer a greater REY reservoir to tap during formation of deep-water phosphorites. The increases of the REY with water depth are significant, 5–10 times increases from the sea surface to 3000 m water depth, the variation depending on the individual REY and location in the global ocean (e.g., [24]). A corollary to this is that the seamount phosphorites formed earlier in the Cenozoic [4], when environmental conditions favored greater supplies of REY from seawater (e.g., [25]), and perhaps slower rates of precipitation of Fe-Mn oxyhydroxides due to intensified suboxic conditions (e.g., [10]). A difficulty in applying inferences from the geological record is that most of the deposits analyzed formed at continental margins and REY concentrations at various water depths through time in open-ocean environments are not as well known.
The association of seamount phosphorites with Fe-Mn oxides is the third fundamental difference with CM phosphorites. However, both groups of phosphorites obtain REY from seawater, which has a REY pattern controlled by complex processes, including the sorption of REY on Fe-Mn oxides and other materials, release of REY from dissolving bio-silica (opal) and bio-calcite, and solution complexation dominated by mono- and di-carbonate complexes [18,23,24,25]. These processes are also water-depth dependent, as with the distribution of REY concentrations, indicating a significant water-depth impact on REY incorporation into the phosphorites. In addition, CM phosphorites obtain REY from pore fluids influenced by diagenetic reactions. Both groups of phosphorites also inherit a small amount of their REY complements from the carbonates that were replaced (e.g., [2]).
The Fe-Mn crusts and phosphorites obtain REY from seawater by different mechanisms, crusts by sorption of the REY and oxidation of the Ce, and phosphorites by substitution for Ca during precipitation, and to lesser extents by inheritance and sorption. Seamount phosphorite mean REY contents normalized to mean Prime Crust Zone (PCZ [26]) Fe-Mn crusts show a very large positive Y anomaly, a smaller positive La anomaly, and smaller yet positive Gd anomaly, as well as a large negative Ce anomaly; this ratio generally increases with increasing atomic number (Figure 6). These anomalies are opposite to the anomalies found for Fe-Mn crusts but are the same anomalies found for seawater [18]. Seawater, Fe-Mn crusts, and phosphorites are HREY enriched relative to LREE, but to different degrees, with seawater at ~2000 m water depth in the central Pacific showing 73% HREY complement of total REY [19] versus a mean of 60% (range 47%–74%) for seamount phosphorites and mean of 18% for PCZ Fe-Mn crusts [27]. The seamount phosphorite REY pattern compares closely with that of intermediate depth seawater in the central Pacific. If crust growth slowed or even stopped during phosphatization of the crusts and substrate rocks, it is possible that this process changed the LREY/HREY ratio and REY concentrations of seawater from which the seamount phosphorites precipitated over million-year time frames. The range of that ratio for the seamount phosphorites is consistent with that idea, but necessary data are not available to verify it.
A fourth major difference is that phosphatization of the seamount deposits occurred in an organic-matter-poor sedimentary environment (e.g., [28]), which did not allow for redox diagenetic zones to be established in the sediment that was cemented and replaced by the CFA. Seamount foraminiferal sands, the most commonly replaced sediment type, and reef limestone [29], have pore waters that are the same as seawater, which during phosphatization were likely suboxic; the upper meters of sediment/limestone act as an open system. In contrast, most CM deposits formed in organic matter-rich environments that promoted complex diagenetic zones that changed as sediment was eroded and deposited. Many chemical reactions in different diagenetic zones, such as Mn and Fe reduction, organic matter oxidation, suboxic diagenesis, anoxic sulfate reduction, methanogenesis, etc., promoted the formation of different authigenic minerals, such as glauconite, pyrite, dolomite, and other metal carbonates (e.g., [11,12,30], and references therein). These diagenetic processes may in part fractionate the REY among the authigenic phases, including the CFA in phosphorite; the diagenetic fluid that contributes to the formation of the phosphorite is less enriched in REY compared to seawater unaltered by these diagenetic reactions.

4.4. Comparisons with Terrestrial Carbonatite REY Deposits and Marine Fe-Mn Deposits

These marine phosphorites contrast with the typical, large, land-based carbonatite-hosted REY, which have a HREY complement of <1%, and very high grades for the LREY [31]. Carbonatites are the deposits from which more than 90% of the REY used by industry are produced. The existing, and soon to be operational, carbonatite-hosted REY deposits can supply the LREY needed by industry, but not the HREY.
Marine Fe-Mn crusts and nodules, which are considered as potential resources for REY, especially the HREY, also have on average much lower HREY complements than the phosphorites, mean 17% for crusts and 22% for nodules (Figure 7); but crusts have total REY grades up to 1%, but more typically 0.20% to 0.40%, while nodules have lower grades, 0.08% for Clarion-Clipperton Zone (CCZ) and 0.17% for Cook island nodules (Figure 8) [27,32,33,34]. As mentioned, it is the HREY complement that is most important because those elements are most needed for high-tech and green-tech applications and have the highest per weight market value (e.g., LCDs, permanent magnets, solar panels, lasers, nuclear medicine, phosphors) [35].
An advantage to extracting REY from phosphorite deposits is that the REY are hosted by the CFA, which is easily dissolved with a mild acid leach [7]. This is not the case for land-based carbonatites, in which the REY are hosted in refractory minerals that are difficult to process. Given this processing advantage, and considering the relatively high grade and %HREY component, seamount phosphorite deposits warrant further research as a potential future resource for REY. An exploration disadvantage for seamount phosphorites however is that they occur in the same places as Fe-Mn crusts on the seamounts and are typically covered by the Fe-Mn crusts. Consequently, it would require the development of new technology to identify phosphorite substrate rock beneath the Fe-Mn crusts, or detailed sampling.
In addition, it is because of this aspect of seamount phosphorites that reliable estimates of tonnages have not been possible to determine whereas tonnages are quite well known for some CM phosphorite deposits. An advantage of Fe-Mn crusts as a source for REY, like phosphorites, is that they also dissolve easily using a weak acid leach, and they are relatively accessible to exploration, although mining technology has yet to be fully developed. For crusts, the REY would be a byproduct of the mining of focus metals, such as Co, Ni, Mo, Mn, and others. The difficult aspect in development of a useful mining technology for crusts is the recovery of the Fe-Mn crust without also collecting substrate rock, which would dilute the grade of the ore. An ideal mine site might be one where the substrate rock is phosphorite. Then it would be desirable to collect both phosphorite and Fe-Mn crusts at the same time, which would provide two potential multi-component ore deposits in one mining operation, one with P and REY co-products and one with focus metals (Co, Ni, Mn) and byproducts (Mo, REY, Te, and others). Byproducts associated with phosphorite recovery are also possible, for example F, especially in seamount phosphorites, and U, V, and others in some CM deposits (Table S1). This would, however, complicate the processing and beneficiation, but is an option to consider during operational planning and exploration.

5. Conclusions

(1)
Continental margin (CM) marine phosphorites have low total REY contents (mean 161 ppm) and high HREY complements (mean 49%), while seamount phosphorites have 4–6-times higher individual REY contents (except for Ce, which is comparable; mean ΣREY 727 ppm), and very high HREY complement (mean 60%).
(2)
The predominant causes of higher concentrations and larger HREY complements in seamount phosphorites compared to CM phosphorites is the geological time of formation, changes in seawater REY concentrations over time, water depth of formation, differences in organic carbon content in the depositional environments and its role in the development of diagenetic zones in the sediment, and possibly the concurrent precipitation of Fe-Mn oxides with the seamount phosphorites.
(3)
Fe-Mn crusts and nodules are another potential resource for REY. Both Fe-Mn deposit types have significantly lower HREY complements than the phosphorites and crusts have three to ten times higher REY concentrations. These differences can be explained by the mechanisms of incorporation of the REY into the mineral deposits: Predominantly sorption by the Fe-Mn oxides with surface oxidation of the Ce, and predominantly by substitution for Ca in the CFA structure, and to lesser extents by inheritance from the host rock and by sorption.
(4)
Seamount phosphorites occur in the same places as Fe-Mn crusts on seamounts and the phosphorites are typically covered with Fe-Mn crusts, making exploration for phosphorites challenging. Consequently, it would require the development of new technology or detailed sampling to identify phosphorite substrate rock beneath the Fe-Mn crusts. Detailed sampling and geophysical measurements of a single guyot might provide geological or morphological criteria that would be useful for future exploration.
(5)
An ideal mine site might be one where the substrate rock is phosphorite, in which case it would be desirable to collect both phosphorite and Fe-Mn crusts at the same time. This approach would provide two potential multi-component ore deposits in one mining operation, one with P and REY co-products and one with focus metals (Co, Ni, and Mn) and byproducts (e.g., Mo, REY, and Te).
(6)
Potential ore deposits with high HREY complements, like the marine phosphorites analyzed in this paper, could help supply the HREY needed for high-tech and green-tech applications without creating an oversupply of the LREY. Consequently, a search for such deposits with the largest HREY/LREY ratios would help alleviate the supply/oversupply problem.
(7)
Land-based phosphorite deposits offer a similar potential REY resource as a byproduct or co-product of the focus phosphate mining (e.g., [6,7]). Recovery of these land-based REY would require the addition of costly infrastructure and changes in extractive processing to the existing mining operations. Production of REY as a co-product of phosphorite mining would be advantageous if considered in the early stages of planning a new mining operation, such as those currently proposed for CM offshore sites.

Supplementary Materials

The following are available online at www.mdpi.com/2075-163X/6/3/88/s1, Table S1: Chemical composition of marine phosphorites, Figure S1: Photographs of continental margin phosphorites: (A) A pervasively Carbonate Fluorapatite (CFA)-replaced breccia, California margin; field of view represents 14 cm; (B) Phosphorite gravel with shell fragments from Chatham Rise, Table S2: Statistics of phosphorite deposits from continental margins, Table S3: Statistics of all phosphorite deposits from seamounts.

Acknowledgments

We thank the captain, ship’s crew, and the scientific crew of the R.V. SP Lee Southern California Borderland cruises, 1974 and 1976, R.V. Gosnold Blake Plateau cruises, 1964 and 1965, Chain 46 Blake Plateau cruise, 1966, R.V. Atlantis II 266 Blake Plateau cruise, 1961, R.V. Seward Johnson and submersible Johnson Sea-Link Peru Margin expedition (SJ 1092), 1992, R.V. Robert D. Conrad Peru Margin cruise (23-06), 1982, R.V. Melville Majuro, Gilbert Islands, and Samoa cruise, 1999, R.V. Farnella Marshall Islands cruise, 1989, R.V. Farnella Hawaii cruise, 1989, R.V. Thomas G. Thompson Shatsky Rise cruise, 1994, R.V. Thomas Washington Northwest Pacific cruise, 1991, the R.V. Vinogradov Hawaii and Johnston Island cruise, 1991, R.V. Sonne 1 Southeast Pacific cruise, 1989/1990, R.V. Sonne 1 North Atlantic cruise, 1992, R.V. SP Lee Horizon Guyot, Necker Ridge, and Johnston Island cruise, 1983, R.V. Farnella Johnston Island and Karin Ridge cruise, 1986, and R.V. Farnella Hawaii cruise, 1988. We thank Ellen Roosen, WHOI and Brian Buczkowski, USGS for supplying Blake Plateau samples. We thank Lydia Somers for preparation and analysis of some of the seamount samples and Kira Arias La Rheir for preparation of samples. We thank the two anonymous reviewers and Amy Gartman, USGS, for useful comments that helped improve this paper.

Author Contributions

James R. Hein., Craig R. Glenn. and Ray Wood were responsible for collection of samples; James R. Hein, Andrea Koschinsky, Mariah Mikesell, and Kira Mizell for analyses of samples and data; and James R. Hein and Andrea Koschinsky for interpretation of data. All authors contributed to the final version of this manuscript.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. Any use of trade, product, or firm names is for descriptive purposes only and does not imply endorsement by the authors or their affiliated institutions.

References

  1. Altschuler, Z.S. The geochemistry of trace elements in marine phosphorites Part I. Characteristic abundances and enrichment. In Marine Phosphorites—Geochemistry, Occurrence, Genesis; Bentor, Y.K., Ed.; Society of Economic Paleontologists and Mineralogists: Tulsa, OK, USA, 1980; pp. 19–30. [Google Scholar]
  2. McArthur, J.M.; Walsh, J.N. Rare-earth geochemistry of phosphorites. Chem. Geol. 1984, 47, 191–220. [Google Scholar] [CrossRef]
  3. Watkins, R.T.; Nathan, Y.; Bremner, J.M. Rare earth elements in phosphorite and associated sediment from the Namibian and South African continental shelves. Mar. Geol. 1995, 129, 111–128. [Google Scholar] [CrossRef]
  4. Hein, J.R.; Yeh, H.-W.; Gunn, S.H.; Sliter, W.V.; Benninger, L.M.; Wang, C.-H. Two major Cenozoic episodes of phosphogenesis recorded in equatorial Pacific seamount deposits. Paleoceanography 1993, 8, 293–311. [Google Scholar] [CrossRef]
  5. Glenn, C.R.; Föllmi, K.B.; Riggs, S.R.; Baturin, G.N.; Grimm, K.A.; Trappe, J.; Abed, A.M.; Galli-Olivier, C.; Garrison, R.E.; Ilyin, A.V.; et al. Phosphorus and phosphorites: Sedimentology and environments of formation. Eclogae Geol. Helv. 1994, 87, 747–788. [Google Scholar]
  6. Christmann, P. A forward look into rare earth supply and demand: A role for sedimentary phosphate deposits? Procedia Eng. 2014, 83, 19–26. [Google Scholar] [CrossRef]
  7. Emsbo, P.; McLaughlin, P.I.; Breit, G.N.; du Bray, E.A.; Koenig, A.E. Rare earth elements in sedimentary phosphate deposits: Solution to the global REE crisis? Gondwana Res. 2015, 27, 776–785. [Google Scholar] [CrossRef]
  8. Santos, A.J.G.; Mazzilli, B.P.; Fávaro, D.I.T.; Silva, P.S.C. Partitioning of radionuclides and trace elements in phosphogypsum and its source material based on sequential extraction methods. J. Environ. Radioact. 2006, 87, 52–61. [Google Scholar] [CrossRef] [PubMed]
  9. Weng, Z.; Jowitt, S.M.; Mudd, G.M.; Haque, N. A detailed assessment of global rare earth element resources: Opportunities and challenges. Econ. Geol. 2015, 110, 1925–1952. [Google Scholar] [CrossRef]
  10. Koschinsky, A.; Stascheit, A.; Bau, M.; Halbach, P. Effects of phosphatization on the geochemical and mineralogical composition of marine ferromanganese crusts. Geochim. Cosmochim. Acta 1997, 61, 4079–4094. [Google Scholar] [CrossRef]
  11. Glenn, C.R.; Arthur, M.A. Petrology and major element geochemistry of Peru margin phosphorites and associated diagenetic minerals: Authigenesis in modern organic-rich sediments. Mar. Geol. 1988, 80, 231–267. [Google Scholar] [CrossRef]
  12. Glenn, C.R.; Arthur, M.A.; Resig, J.M.; Burnett, W.C.; Dean, W.E.; Jahnke, R.A. Are Modern and Ancient Phosphorites Really So Different? In Siliceous, Phosphatic and Glauconitic Sediments of the Tertiary and Mesozoic; Iijima, A., Abed, A.M., Garrison, R.E., Eds.; VSP International Science Publishers: Leiden, The Netherlands, 1994; pp. 159–188. [Google Scholar]
  13. Kudrass, H.; Cullen, D. Submarine phosphorite nodules from the Central Chatham Rise off NZ-composition, distribution and reserves—(Valdivia Cruise 1978). Geol. Jahrb. 1982, D51, 3–41. [Google Scholar]
  14. Cook, H.E.; Johnson, P.D.; Matti, J.C.; Zemmels, I. Methods of sample preparation and X-ray diffraction data analysis (X-ray mineralogy laboratory, Deep Sea Drilling Project, University of California Riverside). In Initial Reports of the Deep Sea Drilling Project; U.S. Government Printing Office: Washington, DC, USA, 1975; Volume 28, pp. 999–1007. [Google Scholar]
  15. Hein, J.R.; Schwab, W.C.; Davis, A.S. Cobalt and platinum-rich ferromanganese crusts and associated substrate rocks from the Marshall Islands. Mar. Geol. 1988, 78, 255–283. [Google Scholar] [CrossRef]
  16. McLennan, S.M. Rare earth elements in sedimentary rocks: Influence of provenance and sedimentary processes. In Geochemistry and Mineralogy of Rare Earth Elements; Lipin, B.R., McKay, G.A., Eds.; Mineralogical Society of America’s Reviews in Mineralogy: Washington, DC, USA, 1989; Volume 21, pp. 169–200. [Google Scholar]
  17. Anders, E.; Grevesse, N. Abundances of the elements: Meteoritic and solar. Geochim. Cosmochim. Acta 1989, 53, 197–214. [Google Scholar] [CrossRef]
  18. 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]
  19. 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]
  20. Hughes, J.M.; Cameron, M.; Mariano, A.N. Rare-earth-element ordering and structural variations in natural rare-earth-bearing apatites. Am. Mineral. 1991, 76, 1165–1173. [Google Scholar]
  21. Hughes, J.M.; Rakovan, J.F. Structurally robust, chemically diverse: Apatite and apatite supergroup minerals. Elements 2015, 11, 165–170. [Google Scholar] [CrossRef]
  22. Piper, D.Z.; Perkins, R.B.; Rowe, H.D. Rare-earth elements in the Permian Phosphoria Formation: Paleo proxies of ocean geochemistry. Deep Sea Res. II 2007, 54, 1396–1413. [Google Scholar] [CrossRef]
  23. De Baar, H.J.W.; Bacon, M.P.; Brewer, P.G.; Bruland, K.W. Rare earth elements in the Pacific and Atlantic Oceans. Geochim. Cosmochim. Acta 1985, 49, 1943–1959. [Google Scholar] [CrossRef]
  24. Zhang, J.; Nozaki, Y. Rare earth element and yttrium in seawater: ICP-MS determination in the East Caroline, Coral Sea, and South Fiji basins of the western South Pacific Ocean. Geochim. Cosmochim. Acta 1996, 60, 4631–4644. [Google Scholar] [CrossRef]
  25. Lécuyer, C.; Reynard, B.; Grandjean, P. Rare earth element evolution of Phanerozoic seawater recorded in biogenic apatites. Chem. Geol. 2004, 204, 63–102. [Google Scholar] [CrossRef]
  26. Hein, J.R.; Conrad, T.A.; Dunham, R.E. Seamount characteristics and mine-site model applied to exploration- and mining-lease-block selection for cobalt-rich ferromanganese crusts. Mar. Georesour. Geotechnol. 2009, 27, 160–176. [Google Scholar] [CrossRef]
  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 deposits. Ore Geol. Rev. 2013, 51, 1–14. [Google Scholar] [CrossRef]
  28. Baturin, G.N. Thallium in oceanic phosphorites. Dokl. Earth Sci. 2004, 394, 68–72. [Google Scholar]
  29. Benninger, L.M.; Hein, J.R. Diagenetic evolution of seamount phosphorite. In Marine Authigenesis: From Global to Microbial; Glenn, C.R., Prévôt-Lucas, L., Lucas, J., Eds.; SEPM Special Publication: Tulsa, OK, USA, 2000; pp. 245–256. [Google Scholar]
  30. Jarvis, I.; Burnett, W.C.; Nathan, Y.; Almbaydin, F.S.M.; Attia, A.K.M.; Castro, L.N.; Flicoteaux, R.; Hilmy, M.E.; Husain, V.; Quatawnah, A.A.; et al. Phosphorite geochemistry: State-of-the-art and environmental concerns. Eclogae Geol. Helv. 1994, 87, 643–700. [Google Scholar]
  31. Long, K.R.; van Gosen, B.S.; Foley, N.K.; Cordier, D. The Principal Rare Earth Elements Deposits of the United States—A Summary of Domestic Deposits and a Global Perspective; U.S. Geological Survey Scientific Investigations Report 2010-5220; U.S. Geological Survey: Reston, VA, USA, 2010; p. 96.
  32. Hein, J.R. Prospects for Rare Earth Elements from Marine Minerals; International Seabed Authority Briefing Paper 02/12; International Seabed Authority: Kingston, Jamaica, 2012; pp. 1–4. [Google Scholar]
  33. Hein, J.R.; Koschinsky, A. Deep-ocean ferromanganese crusts and nodules. In Treatise on Geochemistry, 2nd ed.; Holland, H.D., Turekian, K.K., Eds.; Elsevier: Oxford, UK, 2014; Volume 13, Chapter 11; pp. 273–291. [Google Scholar]
  34. 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]
  35. Stegen, K.S. Heavy rare earths, permanent magnets, and renewable energies: An imminent crisis. Energy Policy 2015, 79, 1–8. [Google Scholar] [CrossRef]
Figure 1. Location of samples used in this study. See Table S1 for coordinates.
Figure 1. Location of samples used in this study. See Table S1 for coordinates.
Minerals 06 00088 g001
Figure 2. Photographs of seamount phosphorites: (A) cement-supported breccia; cement from carbonate fluorapatite (CFA)-replaced pelagic foraminifera matrix. (B) Massive, layered (AD), recrystallized phosphorite; protolith cannot be distinguished, low porosity. (C) Massive bed of CFA-replaced foraminifera, with nearly pristine microfossil structure; minor rim cement on foraminifera indurates the rock; very high porosity. (D) CFA-replaced carbonate sand that filled fractures in basalt. Note Fe-Mn crust on at least one surface of each sample.
Figure 2. Photographs of seamount phosphorites: (A) cement-supported breccia; cement from carbonate fluorapatite (CFA)-replaced pelagic foraminifera matrix. (B) Massive, layered (AD), recrystallized phosphorite; protolith cannot be distinguished, low porosity. (C) Massive bed of CFA-replaced foraminifera, with nearly pristine microfossil structure; minor rim cement on foraminifera indurates the rock; very high porosity. (D) CFA-replaced carbonate sand that filled fractures in basalt. Note Fe-Mn crust on at least one surface of each sample.
Minerals 06 00088 g002
Figure 3. Scatter plot of phosphorus versus calcium for seamount and continental margin phosphorite deposits analyzed here. Arrows show deviations from the carbonate fluorapatite end member (most pure phosphorite) with increasing calcite and detrital/authigenic (Si, Al, and Fe) contents.
Figure 3. Scatter plot of phosphorus versus calcium for seamount and continental margin phosphorite deposits analyzed here. Arrows show deviations from the carbonate fluorapatite end member (most pure phosphorite) with increasing calcite and detrital/authigenic (Si, Al, and Fe) contents.
Minerals 06 00088 g003
Figure 4. Bar diagram of mean contents of individual rare earth elements and yttrium for the sample groups studied here; note that the order for each element follows the order in the key.
Figure 4. Bar diagram of mean contents of individual rare earth elements and yttrium for the sample groups studied here; note that the order for each element follows the order in the key.
Minerals 06 00088 g004
Figure 5. Post Archean Australian Shale-normalized rare earth element plot for mean data of sample groups compared with patterns for Prime Crust Zone (PCZ) crusts, Clarion-Clipperton Zone (CCZ) nodules, and seawater (× 109) at ~2000 m water depth [19].
Figure 5. Post Archean Australian Shale-normalized rare earth element plot for mean data of sample groups compared with patterns for Prime Crust Zone (PCZ) crusts, Clarion-Clipperton Zone (CCZ) nodules, and seawater (× 109) at ~2000 m water depth [19].
Minerals 06 00088 g005
Figure 6. Bar diagram of ratio of mean REY data for seamount phosphorites and Pacific Prime Crust Zone ferromanganese crusts.
Figure 6. Bar diagram of ratio of mean REY data for seamount phosphorites and Pacific Prime Crust Zone ferromanganese crusts.
Minerals 06 00088 g006
Figure 7. Comparison of mean total HREY contents of ferromanganese deposits with seamount and continental margin phosphorites.
Figure 7. Comparison of mean total HREY contents of ferromanganese deposits with seamount and continental margin phosphorites.
Minerals 06 00088 g007
Figure 8. Bar diagram of mean total REY data for ferromanganese crusts and nodules compared to mean data for seamount and continental margin phosphorites.
Figure 8. Bar diagram of mean total REY data for ferromanganese crusts and nodules compared to mean data for seamount and continental margin phosphorites.
Minerals 06 00088 g008
Table 1. XRD mineralogy of phosphorites from seamounts 1.
Table 1. XRD mineralogy of phosphorites from seamounts 1.
Cruise IDSample IDMajor 2ModerateMinor
AVON0211-MassiveCFAcalcite--
AVON0211-PorousCFA--calcite
AVON0255-13CFA--K-feldspar, chlorite
F10-89-CPD10-17CFA----
F7-86-HWCD19-1ACFA--calcite, quartz, plagioclase
F7-86-HWCD30-8CFAplagioclasechlorite
TN037D10-M-1CFA--phillipsite, quartz
TN037D12-2CFA--quartz
TUNES6D32-6CFAplagioclasechlorite, smectite, sepiolite
TUNES6D32-9CFAcalcitebarite
V2-91-CPD4-12CFA--K-feldspar, smectite, amphibole
V2-91-CPD7-8CFA--quartz, ilmenite, palygorskite
Leached Samples 3
SO6626 DSR 1CFA----
SO6626 DSR 2CFA----
SO66 428 DSR 6CFA----
SO6629 DSR 5CFA--phillipsite
SO6650 DSR 3CFA----
SO6661 DSR 1CFA----
SO66 469 DSO 2CFAphillipsite--
SO6680 DSK 6CFA----
SO6685 DSK 1CFA----
L5-83-HW 4D5-A3-2CFA----
L5-83-HW 4D29-A1-bCFA----
F7-86-HW 4CD14-2DCFA----
F2-88-HW 4D12-1ACFA--quartz
F2-88-HW 4D12-5CFA--quartz
SO8364 GTV 1CFA----
SO83118 GTV 6CFA----
SO83126 GTV 3CFA----
1 Major >25%; Moderate 5%–25%; Minor <5%; 2 CFA = carbonate fluorapatite; 3 Samples treated with 0.5 M triammonium citrate and 30% H2O2; 4 Mineralogy of sample did not change after leaching treatment; -- means not present.
Table 2. XRD Mineralogy of phosphorites from continental margins 1.
Table 2. XRD Mineralogy of phosphorites from continental margins 1.
Cruise IDSample IDMajor 2ModerateMinor
CA Borderland
L-1-74-SCLCB-14-ACFAPxanalcime, quartz
L-1-74-SCLCB-14-BCFAPx, plagioclaseanalcime
L-1-74-SCLCB-1-4BCFA--quartz, dolomite
L-1-74-SCLCB-1-4CCFAquartz, plagioclase--
L-1-74-SCLCB-20-6CCFAplagioclasequartz
L-1-74-SCLCB-20-8BCFA--Quartz (?)
L-2-76-SCCD281-ACFA--quartz
L-2-76-SCCD281-BCFA--Plagioclase (?)
L-2-76-SCCD350-ACFA--glauconite or illite, quartz (?)
L-2-76-SCCD350-BCFA--quartz, glauconite or illite
Peru Margin
JSL3350-1aCFAquartz, plagioclase--
JSL3350-1bCFAquartz, plagioclaseIllite (?), amphibole (?)
JSL3352-4CFAplagioclase, quartz--
JSL3355-2CFAquartz, plagioclase--
JSL3363-4CFAquartz, plagioclase--
JSL3365-2CFAquartz--
JSL3372-1CFA, quartzplagioclase, pyrite--
JSL3365CFAplagioclase, quartz--
JSL3352-6CFAplagioclasequartz
RC-23061-2, GS-1CFAplagioclase, Px, quartzK-feldspar, glauconite
Chatham Rise
NADD9+2-8CFAASiglauconite, pyrite (?)
NADD9+8CFA, calcite--plagioclase, glauconite, quartz
NADD16+2-8CFAcalcitequartz, glauconite, pyrite
NADD16+8CFAcalcite, goethitequartz, glauconite
NADD19+2-8CFAcalcite, ASiglauconite, pyrite
NADD19+8CFAcalcite, plagioclasequartz, glauconite
NADD21+2-8CFAASicalcite, glauconite, pyrite (?)
Chatham Rise
NADD21+8CFAcalcitegoethite, pyrite, plagioclase, glauconite (?)
NADD22+2-8CFAcalcitequartz, glauconite, pyrite (?)
NADD22+8CFA, calciteplagioclase--
NADD23+2-8CFAcalcite, ASipyrite, glauconite, quartz (?)
NADD23+8CFA, calcite--glauconite or illite
NADD24+2-8aragonite, calciteCFAplagioclase, glauconite, chlorite (?)
NADD24+8CFA, calcite--illite/smectite (mixed layer) (?)
Blake Plateau
GOS 742397-ACFAquartz, calciteglauconite
GOS 742397-BCFAcalcitequartz, glauconite or illite
GOS 742399-ACFAcalcite, goethiteglauconite or illite
GOS 742399-BCFAcalciteglauconite or illite
GOS 742476-ACFA, calcite--quartz
GOS 742476-BCFA, calcite--k-feldspar, birnessite
GOS 742480CFAcalciteglauconite or illite, quartz, birnessite
GOS 742485CFAcalciteglauconite or illite, quartz
CHN46-1Station 15-2CFAcalcite, 10 Å manganate--
AT266D42-4CFAcalcite, 10 Å manganateanalcime
E. Atlantic 3
SO8323 GTV II A3CFA--quartz
1 Major >25%; Moderate 5%–25%; Minor <5%; 2 CFA = carbonate-fluorapatite; Px = pyroxene; ASi = amorphous silica; 10 Å manganate = todorokite, buserite, or asbolane; 3 Sample treated with 0.5 M triammonium citrate and 30% H2O2; -- means not present; (?) means presence uncertain; NA means not available.
Table 3. Statistics of phosphorite deposits from the California Borderland (see Table S1 for full dataset).
Table 3. Statistics of phosphorite deposits from the California Borderland (see Table S1 for full dataset).
ElementNMean 1MedianStDevMinMax
Fe (wt %)101.601.370.900.503.27
Si103.732.582.800.739.07
Al100.950.610.940.163.01
Si/Al103.945.121.362.686.24
Mg100.710.570.330.431.36
Ca1030.632.13.8322.834.5
Na101.271.180.271.021.85
K100.410.440.190.150.72
Ti100.190.030.330.010.89
P1011.612.31.678.1613.2
Ca/P102.642.630.072.572.79
P2O51026.628.13.8418.730.3
S101.101.130.280.621.58
SO3102.742.810.691.553.93
CO2106.316.600.934.347.06
LOI 21011.111.00.6910.212.2
H2O+104.714.750.234.405.05
As (ppm)1013136.16.027
Ba106846474871451290
Cl1010329053686701790
Co105.84.33.52.813
Cr101111074543181
Cu1023205.01831
F1020,90922,7505205899025,400
Li106.44.55.72.020
Mn1079339117256
Mo10106.7124.044
Ni1044436.03555
Pb104.23.61.72.27.6
Sr102072210025715602340
U10106836762285
V101011154149165
Zn105749194194
Zr103716468.0141
La1023.623.06.3715.132.6
Ce1018.714.711.57.0042.6
Pr104.233.771.782.057.39
Nd1016.815.07.087.9029.4
Sm103.433.031.501.606.10
Eu100.970.820.480.431.85
Gd104.634.491.492.486.53
Tb100.710.700.220.381.00
Dy104.614.661.292.636.34
Y1059.759.713.440.278.4
Ho101.071.100.260.671.44
Er103.513.540.822.254.55
Tm100.490.500.110.310.63
Yb102.792.850.561.903.60
Lu100.430.430.090.300.58
ΣREY 3101461444185205
%HREY 410545764261
Cecn 5100.630.620.140.380.85
Cesn 5100.470.470.100.290.64
1 Mean, Median, Min, Max in wt % (Fe to As) and ppm (As and below); 2 %Loss on ignition at 1000 °C; H2O+ is structural water; 3 ΣREY = Sum of rare earth elements plus yttrium; 4 %HREY (Eu–Lu) of the total sum of REY; 5 Cecn = Ce anomaly using chondrite-normalized values; Cesn = Ce anomaly using PAAS-normalized values.
Table 4. Statistics of phosphorite deposits from Peru Margin.
Table 4. Statistics of phosphorite deposits from Peru Margin.
ElementNMean 1MedianStDevMinMax
Fe (wt %)101.721.550.601.172.97
Si108.477.413.085.0514.6
Al102.001.680.661.343.31
Si/Al104.234.100.503.514.87
Mg100.660.670.090.420.74
Ca1025.226.33.0919.128.7
Na101.311.260.171.051.60
K100.730.650.260.441.21
Ti100.090.080.020.070.13
P109.6910.11.277.2411.1
Ca/P102.602.600.042.532.70
P2O51022.223.12.9116.625.5
S101.401.370.540.732.63
SO3103.503.421.351.826.55
CO2105.906.221.014.037.02
LOI 21010.610.81.508.0713.3
H2O+104.494.550.663.505.40
As (ppm)102926207.081
Ba101361066484281
Cl101829176259510073037
Co103.94.01.02.75.5
Cr1076702349116
Cu1013142.28.415
F1022,15720,58837471817329,394
Li1010112.07.014
Mn1098912171141
Mo101712172.963
Ni104342182181
Pb106.46.01.84.29.8
Sr101756185523813301980
U101151233548177
V106264124080
Zn1040408.33056
Zr107175223695
La1010.410.51.258.5012.6
Ce1017.716.82.9814.723.5
Pr102.302.220.321.962.93
Nd109.099.201.187.5011.1
Sm101.841.850.261.502.20
Eu100.470.460.090.330.62
Gd102.002.090.311.412.46
Tb100.310.310.040.240.37
Dy101.961.980.351.272.39
Y1018.818.44.399.6024.2
Ho100.440.440.080.280.54
Er101.331.340.240.831.70
Tm100.210.210.040.120.25
Yb101.321.330.220.801.60
Lu100.220.220.040.120.27
ΣREY 310687095280
%HREY 410404152945
Cecn 5101.131.120.041.071.19
Cesn 5100.850.840.030.810.90
See Table 3 for footnotes.
Table 5. Statistics of phosphorite deposits from Chatham Rise.
Table 5. Statistics of phosphorite deposits from Chatham Rise.
ElementNMean 1MedianStDevMinMax
Fe (wt %)153.243.081.231.265.00
Si154.284.621.911.336.55
Al150.580.600.310.171.39
Si/Al157.417.651.953.9311.4
Mg150.620.590.220.340.93
Ca1531.030.93.0827.035.7
Na150.700.690.040.630.78
K151.051.010.590.331.92
Ti15<0.04<0.010.09<0.010.31
P158.368.771.505.459.86
Ca/P153.723.650.992.755.73
P2O51519.220.13.4312.522.6
S150.640.640.120.470.90
SO3151.591.600.291.172.25
CO225.825.820.435.516.12
LOI 21516.916.85.229.9724.2
H2O+24.054.050.074.004.10
As (ppm)1522237.91139
Ba1594509134335
Cl158508103463401630
Co156.77.72.13.09.7
Cr154044201576
Cu158.07.04.51.017
F1517,86017,400408511,00024,800
Li157.78.03.63.013
Mn1593786557324
Mo152.92.81.50.947.3
Ni153030101553
Pb1512124.75.220
Sr151447133049311102660
U151551307764334
V157173164794
Zn1527275.41738
Zr1524229.71052
La1522.720.413.67.6048.9
Ce1515.615.88.054.1029.8
Pr152.842.751.351.025.38
Nd1511.511.25.414.3021.7
Sm151.971.900.860.803.60
Eu150.490.470.220.180.90
Gd152.712.521.271.095.11
Tb150.390.360.180.160.73
Dy152.542.321.210.984.88
Y1537.633.821.013.678.1
Ho150.670.610.340.251.35
Er152.041.841.010.824.09
Tm150.290.260.140.120.57
Yb151.751.600.840.703.40
Lu150.260.220.130.100.53
ΣREY 315103955536209
%HREY 415474833850
Cecn 5150.780.810.080.590.87
Cesn 5150.590.610.060.450.66
See Table 3 for footnotes.
Table 6. Statistics of phosphorite deposits from Blake Plateau.
Table 6. Statistics of phosphorite deposits from Blake Plateau.
ElementNMean 1MedianStDevMinMax
Fe (wt %)102.352.341.120.803.88
Si102.862.421.511.176.45
Al100.710.750.280.311.28
Si/Al104.044.352.151.607.91
Mg100.650.640.160.430.92
Ca1031.832.22.0727.134.9
Na100.710.700.120.530.91
K100.510.500.330.191.27
Ti100.050.040.030.020.11
P108.808.841.256.7610.9
Ca/P103.613.530.722.935.17
P2O51020.220.32.8615.525.0
S100.680.660.140.510.96
SO3101.701.640.351.272.40
CO21013.012.94.016.6220.0
LOI 21016.716.43.8411.323.8
H2O+103.703.900.702.454.50
As (ppm)103534151259
Ba10323217394221301
Cl10506457202199895
Co10126561375.6338
Cr1099829212269
Cu107217955.4290
F1027,21627,516324821,91232,171
Li1014127.77.031
Mn109108464711,56523230,051
Mo102213223.572
Ni10467107583261738
Pb101913144.940
Sr101376135430010011990
U106236845.3285
V101131223937174
Zn1087934139160
Zr1073794322151
La1071.157.942.523.4133
Ce1036.234.420.77.6077.8
Pr1010.88.486.733.8322.6
Nd1048.037.130.617.0102
Sm109.097.105.933.1020.1
Eu102.451.931.570.885.23
Gd1012.910.18.434.4827.1
Tb101.711.321.100.613.53
Dy1011.88.927.604.1723.1
Y1013010183.738.4257
Ho102.581.971.660.874.76
Er107.755.804.972.5514.8
Tm101.170.880.750.382.32
Yb106.654.904.362.2013.9
Lu101.190.870.790.382.52
ΣREY 310353286211124663
%HREY 410505173359
Cecn 5100.550.500.290.271.24
Cesn 5100.410.380.220.200.94
See Table 3 for footnotes.
Table 7. Statistics of unleached phosphorite deposits from seamounts.
Table 7. Statistics of unleached phosphorite deposits from seamounts.
ElementNMean 1MedianStDevMinMax
Fe (wt %)120.660.420.750.062.89
Si122.331.862.110.037.15
Al120.780.640.650.022.12
Si/Al122.992.870.521.553.37
Mg120.270.240.140.110.62
Ca1234.735.43.2526.237.8
Na120.800.830.150.561.09
K120.360.270.310.020.86
Ti12<0.08<0.050.08<0.010.28
P1212.712.91.279.9514.0
Ca/P122.732.660.262.583.51
P2O51229.129.72.9022.832.2
S120.680.710.090.540.80
SO3121.711.760.221.352.00
CO295.125.131.223.756.96
LOI 2129.148.082.906.6717.5
H2O+92.222.100.321.902.80
As (ppm)125.25.01.73.09.0
Ba1248381725201900
Cl1212087559583103410
Co122112221.767.8
Cr122114157.051.0
Cu1277645615205
F1225,76728,250546112,30030,300
Li12<6.6<6.05.9<1.020
Mn1212107751343463873
Mo123.12.82.30.757.3
Ni121166111521366
Pb127.06.35.11.115
Sr12122311753387241920
U126.76.31.94.711
V123333102259
Zn1251492026100
Zr129179529.5211
La1210279.761.816.6214
Ce1215.013.17.563.1030.2
Pr1214.49.2110.32.8536.8
Nd1260.638.843.611.9153
Sm1211.57.058.822.4031.1
Eu123.262.162.430.738.31
Gd1218.213.212.83.4742.5
Tb122.441.871.650.465.67
Dy1217.414.511.33.0338.7
Y1222420513238.9437
Ho124.193.752.540.718.69
Er1213.412.37.882.2926.9
Tm121.881.771.100.343.88
Yb1212.912.37.622.4027.2
Lu121.981.921.190.354.19
ΣREY 3125034342921121022
%HREY 412605884774
Cecn 5120.250.170.340.061.29
Cesn 5120.190.130.250.050.98
See Table 3 for footnotes.
Table 8. Statistics of leached phosphorite deposits from seamounts a.
Table 8. Statistics of leached phosphorite deposits from seamounts a.
ElementNMean 1MedianStDevMinMax
Fe (wt %)180.460.390.390.011.38
Si181.671.261.440.325.28
Al180.580.430.540.051.88
Si/Al182.883.131.102.196.01
Mg180.210.200.110.040.45
Ca1835.436.32.0830.437.9
Na180.660.660.110.500.85
K18<0.17<0.060.19<0.010.69
Ti18<0.05<0.040.050.0040.22
P1813.513.80.8411.814.6
Ca/P182.622.630.042.582.77
P2O51831.031.61.9327.033.5
S180.710.720.150.490.94
SO3181.771.790.371.222.35
CO2186.356.401.353.847.99
LOI 2188.738.581.445.8811.7
H2O+52.142.340.331.762.45
As (ppm) b0----------
Ba1344012675211952140
Cl18497400646502980
Co1311105.34.023
Cr1318187.19.033
Cu153834201589
F1839,01139,250407831,80046,000
Li135.55.02.72.010
Mn187093001550756816
Mo125.55.52.21.09.0
Ni132321159.071
Pb131511464191218
Sr151563141045310002580
U16128.57.36.330
V133535102150
Zn1380634938205
Zr134438182077
La1818018989.27.20372
Ce1828.529.417.92.0074.0
Pr1823.120.013.71.0049.3
Nd1810288.960.24.40222
Sm1820.317.312.90.9746.8
Eu185.334.773.210.2211.8
Gd1832.229.919.11.4073.9
Tb184.694.452.800.1810.7
Dy1830.829.617.81.4069.7
Y13441415233145976
Ho188.168.384.640.4119.2
Er1824.625.313.81.2057.3
Tm183.513.501.990.208.30
Yb1821.720.812.41.3052.2
Lu133.743.412.001.038.01
ΣREY 3139358244902501992
%HREY 413626255670
Cecn 5180.200.190.080.050.35
Cesn 5180.150.140.060.040.27
a Samples treated with 0.5 M triammonium citrate and 30% H2O2; b -- = data not analyzed; See Table 3 for other footnotes.

Share and Cite

MDPI and ACS Style

Hein, J.R.; Koschinsky, A.; Mikesell, M.; Mizell, K.; Glenn, C.R.; Wood, R. Marine Phosphorites as Potential Resources for Heavy Rare Earth Elements and Yttrium. Minerals 2016, 6, 88. https://doi.org/10.3390/min6030088

AMA Style

Hein JR, Koschinsky A, Mikesell M, Mizell K, Glenn CR, Wood R. Marine Phosphorites as Potential Resources for Heavy Rare Earth Elements and Yttrium. Minerals. 2016; 6(3):88. https://doi.org/10.3390/min6030088

Chicago/Turabian Style

Hein, James R., Andrea Koschinsky, Mariah Mikesell, Kira Mizell, Craig R. Glenn, and Ray Wood. 2016. "Marine Phosphorites as Potential Resources for Heavy Rare Earth Elements and Yttrium" Minerals 6, no. 3: 88. https://doi.org/10.3390/min6030088

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