C Selective Recovery of Rare Earth Elements From Coal Fly Ash

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

Cite This: Environ. Sci. Technol. XXXX, XXX, XXX−XXX pubs.acs.org/est

Selective Recovery of Rare Earth Elements from Coal Fly Ash


Leachates Using Liquid Membrane Processes
Ryan C. Smith,*,† Ross K. Taggart,† James C. Hower,‡ Mark R. Wiesner,†
and Heileen Hsu-Kim*,†

Department of Civil and Environmental Engineering, Duke University, Durham, North Carolina 27708, United States

Center for Applied Energy Research, University of Kentucky, Lexington, Kentucky 40511, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Coal combustion residues and other geological


Downloaded via UNIV AUTONOMA DE COAHUILA on April 6, 2019 at 16:15:07 (UTC).

waste materials have been proposed as a resource for rare


earth elements (REEs, herein defined as the 14 stable
lanthanides, yttrium, and scandium). The extraction of REEs
from residues often generate acidified leachates that require
highly selective separation methods to recover the REEs from
other major soluble ions in the leachates. Here, we studied two
liquid membrane processes (liquid emulsion membranes,
LEM, and supported liquid membranes, SLM) and compared
them to standard solvent extraction techniques for selective
recovery and concentration of REEs from a leachate of coal fly
ash. All separation methods involved an organic solution of
di(2-ethylhexyl)phosphoric acid dissolved in kerosene or
mineral oil and an acid strippant solution of 5 M nitric acid for the liquid-based separations. The LEM configuration, which
separated REEs by immersing an acid-in-oil emulsion in the ash leachate, resulted in similar recovery percentages of individual
REEs as the conventional solvent extraction approach. The recovery of REEs in the SLM configuration, which involved the
impregnation of the solvent in a hydrophobic membrane, was slower than the LEM process. However, the SLM process was
notably more selective for the heavy (and higher value) REEs, while the conventional extraction and LEM processes were more
selective for the light REEs. A flux-based model of the extraction processes suggested that recovery rates were limited by REE
affinity for the solvent chelator in the SLM, while the rates of REEs separation via LEM were limited by diffusive mass transfer
across the liquid membrane. Altogether, these results help to identify specific steps in the recovery process that future work
should target in the development of scalable liquid membrane separations for REE recovery.

■ INTRODUCTION
Rare earth elements (REEs) are considered critical materials for
and other industrial and construction applications, and the
remaining portion being disposed of in ash ponds and landfills.22
a variety of modern technologies including the consumer Technologies for the extraction and recovery of REEs from
electronics, automotive, energy, and defense industries.1−7 With coal fly ash are not well developed, and this knowledge gap
the history of an unstable REE global supply market as well as creates a need to understand the feasibility of this alternative
environmental concerns of conventional REE mining practices, source of REEs. The first step in extraction typically produces a
there exists a need to develop alternate REE sources and complex acidic leachate containing total REEs at 0.5−5 mg L−1,
improve upon existing methods of REE production.8−10 with individual REE concentrations ranging from 0.001 to 1 mg
Recently, researchers have begun investigating the potential L−1. The leachate mixture will also comprise other major solutes
for coal fly ash, coal residuals, geological and marine brines, and including ions of sodium, silicon, aluminum, iron, and calcium in
other waste residues as alternative resources for REEs.11−20 For the range of 100−1000 mg·L−1 in the acid mixture (e.g., typically
example, previous research by our group has shown that the REE diluted HCl or HNO3). Other potential “low grade” feedstocks
content of coal fly ash is typically greater than average crustal (e.g., leachates of mine wastes, geological brines) have similar
earth values; however, the enrichment depends on the type of levels of matrix complexity.11−20 As such, the 102- to 105-fold
coal feedstock used to generate the ash. Coal fly ashes from difference in concentrations between individual REE and major
Appalachian Basin coals were typically greater in REE content
(average 600 mg of REEs per kg ash) relative to fly ashes of other Received: January 24, 2019
major coal producing regions in the U.S.21 Within the United Revised: March 22, 2019
States, over 97 Mt of coal combustion products are created each Accepted: March 25, 2019
year with greater than half being utilized for concrete, drywall, Published: March 25, 2019

© XXXX American Chemical Society A DOI: 10.1021/acs.est.9b00539


Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

metal ions poses a considerable challenge for economical following major metal element composition 54.1% as SiO2,
recovery of REEs. 28.4% Al2O3, 10.9% Fe2O3, and 1.28% Ca (sample ID no. for this
Here, we present work on two different separation processes ash is 93938 App-FA1 in previous studies describing REE
that use liquid membranes to recover REEs from ash leachate: chemical speciation and leaching methods).21,32,33
liquid emulsion membranes (LEMs) and supported liquid Leachates containing the REEs were produced from this ash
membranes (SLMs; Supporting Information Figure S1). Liquid by alkaline roasting followed by acid leaching, a method that is
membrane processes use nonaqueous solvents with dissolved similar to the previous report.21 In brief, 4 g of fly ash and 4 g of
chelating agents to physically separate the aqueous leachate and powdered sodium hydroxide were mixed in a zirconium crucible
acidic strippant phases while simultaneously recovering REEs. and heated at 450 °C for 30 min in a muffle furnace. The roasted
The aqueous strippant may be stabilized within an emulsion ash was cooled to room temperature and mixed with 1 L of 0.11
(LEM process), or the nonaqueous solvent may be impregnated M HNO3 for 12 h on a stir plate. After mixing, solids were
in a physical membrane (SLM process).23−26 These processes allowed to settle for 1 h and the supernatant was decanted from
differ from standard solvent extraction as both REE extraction the remaining solids. For all separation experiments, the
and recovery occur in the same step. Moreover, the advantage of decanted leachate was used within 8 h of preparation.
liquid membrane process configurations is they maximize In addition to the coal ash leachates, the LEM and SLM
interfacial area of REE mass transfer while also reducing the separations were tested on a synthetic leachate solution that was
required volumes of nonaqueous solvent and chelating agent. formulated based on the average composition of leachate
Previous work by others has demonstrated recovery of solutions generated from four different Appalachian coal fly
neodymium dissolved from rare earth magnets using supported ashes (Table S1 in the Supporting Information). The use of a
liquid membranes.27 However, to our best knowledge no studies synthetic solution enabled consistent testing of the methods, as
have applied this separation approach for coal ash leachates or the alkaline roasting and acid leaching process resulted in some
similar low-grade feedstocks of geological origin. For liquid variation of leachate composition. The synthetic leachate was
membrane processes, kerosene or other organic liquids are prepared by mixing the hydrated nitrate salts of major metals
normally used for the nonaqueous phase.28,29 However, working
(Na, Mg, Ca, Al, and Fe ranging from 10 to 6000 mg L−1) and
with kerosene presents a hazard and handling issue due to its
sodium metasilicate nonahydrate in a solution of HNO3 (5% v/
flammability such that alternative solvents may be attractive.
v). Seven REEs (Sc, Y, Nd, Eu, Tb, Dy, and Er) were added from
The efficacy of less hazardous solvents and cost trade-offs must
their respective standards (Inorganic Ventures) to a final
be evaluated.
The goal of the research was to investigate the application of concentration of 150 μg L−1 each.12 This concentration value is
liquid membrane processes for recovering REEs from leachates approximately 1−100 times greater than concentrations in fly
of coal fly ash as a representative nontraditional REE resource. ash leachates. These REEs were selected for initial testing based
The research examined recovery potential as well as selectivity on their criticality34 and also to span a spectrum of light,
for individual REEs. Recoveries of individual REEs were also medium, and heavy REEs. A comparison of individual major
compared as a function of LEM and SLM process variables such metal and REE concentrations in the real and synthetic leachates
as separation time or the type of solvent (kerosene or mineral is provided Table S1. Unless otherwise noted, all separation
oil). Finally, a mathematical model of mass transfer potential in experiments were performed with real coal ash leachates, as
these systems was developed to better understand rate-limiting described above.
process for REE recovery selectivity. Liquid Emulsion Membrane. The LEM process was based


on methods by others23,25,35,36 and entailed first a preparation of
MATERIALS AND METHODS a water-in-oil emulsion. The oil-phase comprised of 10% (v/v)
DEHPA and 1% or 3% (v/v) Span 80 in either kerosene or
Kerosene (purum) or mineral oil (white, light) were compared
as the nonaqueous phases for the separation experiments. mineral oil with a total final oil-phase volume of 50 mL. This oil
Kerosene is often used at the industrial scale for liquid−liquid phase was combined with 50 mL of 5 M HNO3 stripping
separations. Mineral oil presents a reduced hazard of solution, added dropwise under high speed mixing at 7000 rpm
flammability due in part to its lower volatility. Di-2-ethyl- (T18 Ultra-Turrax disperser, IKA Works, Inc.) for a total
hexylphosphoric acid (DEHPA, 97% purity) was dissolved to a volume of 100 mL for the prepared emulsion. Immediately after
concentration of 10% (v/v) into either kerosene or mineral oil preparation, the emulsion was mixed with 500 mL of coal ash
solvent. DEHPA is a strong metal chelator for REEs that is leachate and stirred with a magnetic stir bar at 200 rpm for 15
frequently used for industrial metal separations and is min to 1 h extraction time. After extraction, the leachate/
considered a less hazardous metal carrier than other industrially emulsion mixture was poured into a 2 L glass separatory funnel,
relevant organophosphorus compounds such as tributyl held static for 2 h, and decanted into separate aqueous and
phosphate.30,31 The acid strippant solution for all separations emulsion phases. After emulsion separation into oil and acid
comprised of 5 M HNO3 (≥99.999% purity, trace metal basis). phases, the acid was analyzed for REE and major ion
For the liquid emulsion membrane, Span 80, a hydrophobic concentrations by inductively coupled plasma mass spectrom-
surfactant, was used as a stabilizer for the emulsion. All chemicals etry (ICP-MS), as described below.
were purchased from Sigma-Aldrich and used without further Initial work was first performed using the synthetic leachate to
purification. determine the optimum amount of Span 80 for the kerosene and
Fly Ash Leachate. Leachates were produced from a fly ash mineral oil emulsions. After the surfactant amount was
sample collected in 2014 from a storage silo of a coal-fired power determined, three separate experiments were performed using
plant in Kentucky. This ash sample was produced from a coal a real ash leachate and mineral oil as the diluent for DEHPA and
feedstock originating from the Central Appalachian Basin Span 80 (2%). Due to the difference in volume between acid
region. In a previous study,32 we report that this sample strippant and leachate, the total mass of recovered REEs was
contained total REE content of 703 mg kg−1 and comprised the compared to the total mass of initial REEs in the leachate.
B DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Figure 1. (A) Average percent recovery (n = 3) of major ions and REEs using solvent extraction with 10% DEHPA (v/v) in kerosene after 24 h
extraction. Error bars indicate standard deviation. (B) Mass recovery via liquid emulsion membrane (at 15, 30, and 60 min mixing time) utilizing liquid
phases of 10% DEHPA and 2% Span 80 in mineral oil, and 5 M HNO3 as the acid strippant. (C) Average percent recovery (n = 3) of major elements
and REEs via supported liquid membrane (SLM) using 10% DEHPA in kerosene from a leachate of coal fly ash after 24 h. Error bars indicate standard
deviation. Bar colors correspond to major ions (blue), REEs (green), and radionuclides (red).

Supported Liquid Membrane. A 10% (v/v) solution of The concentrations of the triplicate separations were averaged
DEHPA in kerosene or mineral oil was prepared and passed for each element.
through a 47 mm 0.22 μm PVDF membrane (EMD Millipore) Conventional Solvent Extraction. In addition to the two
by vacuum filtration until the DEHPA solution had soaked liquid membrane processes, standard solvent extraction was also
through the entire membrane. PVDF was chosen for the performed in triplicate using both kerosene and mineral oil at
membrane due to its high hydrophobicity and use by other the same ratios of solvent, acid, and coal ash leachate used for the
researchers.37 The membrane was removed from the vacuum liquid emulsion experiments. A 100 mL aliquot of real coal ash
filtration apparatus and soaked in the remaining 10% DEHPA leachate was separately mixed with 10 mL of 10% DEHPA in
mixture at least 12 h. Then the membrane was placed in an H- mineral oil for 24 h. The organic and aqueous portions were
cell reactor system (Item # MFC 250.40.0, Adams & Chittenden decanted, and 4 mL of mineral oil was mixed with 4 mL of 5 M
Scientific Glass) consisting of two 250 mL flanged glass jars that HNO3 for another 24 h and the acid was analyzed for REE
allowed placement of the prepared 47 mm membrane between content by ICP-MS.
the two glass chambers (see Figure S2 for a schematic). Ash Elemental Analyses. Major and trace element concen-
leachate and 5 M HNO3 stripping solution (250 mL each) were trations were analyzed in the leachate feedstocks and separation
added to opposite sides of the H-cell reactor and stirred at 700 products by ICP-MS (7900, Agilent Technologies). Elements
rpm. Initial testing with synthetic leachate focused on Ca and Si were quantified under hydrogen reaction gas
determining the optimal amount of time to perform extraction, conditions, while all other elements, including the REEs, were
and the effect on recovery due to the difference between using quantified under helium reaction gas mode. Instrument
kerosene and mineral oil as the support phase for DEHPA. For calibrations were performed at the start of a sample batch, and
testing with real leachate, only kerosene was used as the diluent internal standards of Rh and In were spiked in each sample and
for DEHPA. Triplicate SLM extractions were performed on used to correct for shifts in signal intensity during a batch run.
replicate aliquots of a fly ash leachate over 24 h; initial studies on Barium oxides are known to contribute to the mass signal for
extraction demonstrated equilibrium was reached after 20 h europium,38,39 and silicon oxides contribute to the mass signal
(Figure S3). The strippant concentration after 24 h was analyzed for scandium.40 Due to the high concentrations of both barium
for major and trace elements via ICP-MS, as described below. and silicon in ash digests, it is necessary to account for these
C DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Table 1. Estimated Composition, on a Dry Mass Basis, for Major Elements and Total REEs in Initial Leachate and Final Strippant
Products of Liquid Emulsion Membrane (LEM), Supported Liquid Membrane (SLM), and Conventional Solvent Extractiona
15 min LEM 30 min LEM 60 min LEM SLM Kerosene Solvent
initial final initial final initial final initial final initial final
leachate strippant leachate strippant leachate strippant leachate strippant leachate strippant
(mg kg−1) (mg kg−1) (mg kg−1) (mg kg−1) (mg· g−1) (mg kg−1) (mg kg−1) (mg kg−1) (mg kg−1) (mg kg−1)
Na 681 000 410 500 754 000 429 200 796 000 402 600 821 000 702 200 753 000 2079
Mg 7500 18 050 3840 8794 7790 16 600 7450 6297 7850 260
Al 132 000 376 700 103 000 442 300 73 700 385 400 57 000 95 200 99 000 966 800
Si 138 000 40 810 108 000 25 700 97 400 32 330 88 600 124 300 103 000 1574
K 11 700 8396 12 300 8056 13 000 8301 14 200 20 870 12 900 176
Ca 9210 130 200 5780 75 570 7920 143 000 7410 26 960 8870 23 270
Fe 19 100 3054 12 000 2660 2480 739 3050 16 220 12 900 1209
REE 302 3300 146 1710 176 3700 136 5280 213 1820
a
Concentrations were calculated by taking the measured aqueous concentrations and assuming the metals formed nitrate salts upon drying. See the
Supporting Information spreadsheet file.

interferences that affect quantitation of Eu and Sc. A detailed substantial mass recovery of Al and Ca was observed (64% and
explanation of the method for correction can be found in the 17%, respectively). However, radioactive elements such as Th or
Supporting Information, but in brief, the percent oxide U were not recovered to appreciable extents. Note that the
formation for Ba and Si was calculated and then used to correct percentage for each element refers to the final element mass in
the signal intensity at 153 and 45 m/z (mass/charge) for Eu and the acid stipping solution normalized to the element mass in the
Sc, respectively. initial feed solution (e.g., the ash leachate). Concentrations
Separation Factor for Individual Metals. For each values for major ions and REEs for the feed and the strippant
process, the relative separation efficiency for each REE was product solutions are provided in the SI .xlsx data file.
quantified by computing individual distribution coefficients for While the percentage recovery provides information for
each element and comparing this distribution to a reference individual elements, the purity of the mixed REE product is
metal. Here, we defined the distribution coefficient of a easier to compare to REE extraction technologies by calculating
particular metal, D Me , as the ratio between the final a dry mass-based concentration. To estimate this concentration,
concentration in the strippant, [Me]strippant, over the final we used the measured aqueous concentrations of all measured
concentration remaining in the leachate after extraction, metal and metalloid elements (REEs and major cations) in the
[Me]leachate: strippant and determined the dry mass assuming all metal
[Me]strippant cations formed nitrate salts upon drying (e.g., La(NO3)3). Si and
DMe = U were assumed to dry as H4SiO4 and UO2(NO3)2, respectively.
[Me]leachate (1) For the solvent extraction data, the mixed REE purity was
Previous work by Kim et. al using the SLM process to recover estimated to be 1820 mg·kg−1 (dry mass basis), an 8-fold
REEs from scrap magnets using TODGA, showed distribution increase compared to the initial leachate (Table 1).
coefficients for Nd on the order of 103.27 Due to the complex Liquid Emulsion Membrane Processes. The LEM
nature of the coal ash leachate, we compared the relative separation process was first tested on a synthetic leachate to
selectivity of a particular metal to a reference metal, sodium establish surfactant concentrations necessary to form a stable
(Na), by calculating an apparent separation factor, αMe/Na emulsion while also maximize REE separation. We observed a
according to decrease in REE recovery when the surfactant concentration
increased from 1% to 3% in mineral oil (Figure S4). An excess of
DMe [Me]strippant [Na]leachate surfactant at the oil−water interface might hinder transport of
αMe/Na = =
D Na [Na]strippant [Me]leachate (2)
REEs into the oil phase, as indicated by others.23 We also
observed that increasing the concentration of Span 80 improved
Values for αMe/Na were calculated for each REE as well as for the stability of the emulsion. LEM separations using less than 2%
major ions (Ca, Fe, and Al). These calculations utilized the Span 80 did not yield an emulsion that could be stable over for a
following resulting data for each process: (1) the average of 60 min reaction and 2 h settling time. Therefore, 2% Span 80 was
triplicate extractions in the conventional solvent, (2) the 30 min selected for all subsequent experiments with real ash leachates as
mixing time (n = 1) for the LEM process, and (3) the average (n a means to maximize both emulsion stability and REE recovery.
= 3) at the 24 h time point for the SLM process. Na was used as The recovery percentages of REE from the coal ash leachates
the reference because this metal was the greatest concentration by the LEM process were similar to results of the conventional
for all dissolved metals in the ash leachate feedstock. solvent extraction process (Figure 1B). We note that the LEM

■ RESULTS AND DISCUSSION


Conventional Solvent Extraction. The separation and
separation utilized mineral oil as the organic solvent for the
DEHPA (10% v/v) instead of kerosene with DEHPA.
Regardless, the general trend of greater extraction of the light
recovery of REEs from the leachates via conventional solvent REEs over the heavy REEs was observed in the LEM process.
extraction (with kerosene and DEHPA) varied with the type of The mixing time of the LEM process varied between 15, 30,
REE. The recovery percentages for the lighter REEs (La, Pr, Nd, and 60 min with a constant phase separation time of 2 h.
Sm, and Eu) were generally greater than heavier REEs (Dy, Ho, Between mixing times for the LEM process, a 60 min extraction
Er, Tm, Yb, and Lu; Figure 1A). In addition to the REEs, time achieved greater than 70% recovery of most REEs, and as
D DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

the mixing time increased individual recoveries approached the Difficulties with recovering Sc after extraction with DEHPA
recoveries observed with solvent extraction with kerosene, have been known by others for almost 50 years.42 If these
Figure 1B. For all REEs, there was little difference between 15 processes are to be scaled up, economical solutions must be
and 30 min extraction, but a large difference between 30 and 60 developed for recovering Sc and other heavy REEs from the
min extraction. This difference is especially true for most heavy extractant in order to prevent eventual poisoning of the DEHPA.
elements such as Tb, Dy, and Ho that increased from 60%−66% Separation Factors for REE and Major Metal Ions. In
recovery at 15 min extraction to 77%−83% recovery for 60 min comparing the three processes, we observed two important
extraction. Interestingly, the recoveries of Ca and Al switched distinctions: selectivity trends between individual REEs and
between solvent extraction and LEM with Al being recovered selectivity of the REE relative to other competing metals such as
more effectively using kerosene solvent extraction and Ca Ca, Fe, or Al. The apparent separation factors αMe/Na of each
recoveries relatively greater for the mineral oil LEM. REE in the kerosene solvent extraction were always greater than
The expected dry mass REE contents of the LEM product respective αMe/Na values for either liquid membrane processes
were approximately 3300, 1710, and 3700 mg kg−1 for 15, 30, due to the low recovery of Na during solvent extraction (Figure
and 60 min mixing times (Table 1). These values represent a 10-, 2A). The αMe/Na values for the solvent extraction also tended to
12-, and 21-fold increase in purity relative to the initial feed REE
concentrations. For the 15 and 60 min mixing times, REE purity
was almost double the dry mass purity for the kerosene solvent
extraction (1820 mg kg−1.).
Supported Liquid Membrane Process. Initial studies
were performed with the synthetic leachate to understand the
time scale of the SLM process. The results (Figure S3) showed
that the concentrations of REEs in the acid strippant increased
within the first 3 h of separation and, after this period, did not
change considerably between the 3 and 20 h. Based on these
results, we used the 24 h time point to represent an approximate
steady state for subsequent SLM experiments with real coal ash
leachate solutions.
The application of the SLM system for the coal fly ash
leachates resulted in relatively higher percentage recoveries for
the heavy REEs compared to the light REEs (Figure 1C). For
example, the recoveries of Y, Tb, Dy, Ho, Er, Tm, Yb, and Lu
were all greater than 75% after 24 h of extraction, while recovery
of La, Ce, Pr, and Nd was less than 50%. This general difference
between the light and heavy REEs is opposite to the trend
observed for the LEM and conventional solvent extraction
separations. For the major metal constituents Na, Mg, Al, Si, K,
Ca, and Fe, less than 10% was recovered in the strippant
solution. When comparing the expected solid phase product
between the all recovery processes studied (Table 1), the SLM
process resulted in the highest purity with 5280 mg kg−1 total
REEs. The expected solid phase REE content of the original
leachate was only 136 mg kg−1 indicating that the SLM process Figure 2. Selectivity values for kerosene solvent extraction, LEM
was capable of increasing REE content by approximately 39 process, and SLM process for (top) all REEs and (bottom) major
competing metals and light and heavy REEs.
times. Between all three studied processes, the SLM process
provided the highest purity mixed REE product.
However, in addition to the desired REEs, high recovery of decrease slightly with increasing hydrated ionic radius for the
uranium was also observed. Uranium is commonly extracted REE. The LEM αMe/Na values remained fairly constant with
during nuclear fuel reprocessing using tributyl phosphate which increasing hydrated ionic radius apart from Ce, Yb, and Lu while
is similar to the extractant DEHPA used in this study.31 While αMe/Na values for the SLM process increased with ionic radius.
the total recovery was 75%, the expected amount of uranium in In comparisons of the REEs to major competing metals,
the final dried product only would be 0.39% (see the Supporting different trends emerged where the LEM and SLM processes
Information). always show greater selectivity for REEs when compared with
Low Recovery of Scandium. For all three liquid separation solvent extraction, Figure 2B. All three processes demonstrated
processes, Sc was not observed in appreciable concentrations in greater selectivity for REEs over Ca and Fe, but αMe/Na values for
the final acid strippant solution. For example, we observed that Al during solvent extraction were within the same order of
in the SLM process the Sc concentration in the leachate feed magnitude as the light REE Nd. In total, while the αMe/Na values
decreased from an average 29 to 19 μg/L, indicating that 32% of for REEs during solvent extraction were consistently higher than
the original Sc was lost from the leachate but only 1% was both the LEM and SLM processes, compared to other major
recovered into the strippant. The Sc complex with DEHPA has a competing metals like Al, the LEM and SLM processes
very high stability constant (relative to the other REEs). For demonstrated higher selectivity for REEs
example, the estimated binding constant K for the reaction Me3+ Difference between Kerosene and Mineral Oil. The
+ 3 DEHPA2 + Me(DEHP·DEHPA)3 + 3H+ is 1.1 × 107 for Sc3+ importance of the carrier organic solvent for the liquid
(see the Supporting Information) and 1.8 × 103 for Dy3+.41 membrane processes was further explored by directly comparing
E DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Figure 3. Effect of solvent on recovery of selected elements using synthetic leachate for (left) liquid emulsion membrane process with 30 min mixing
time (n = 1) and (right) supported liquid membrane processes after 20 h equilibration time (n = 1). Note that REE separation from real leachate with
the mineral oil solvent was not reproducible.

kerosene and mineral oil for separation of REE in synthetic coal


ash leachates. In the LEM process, the recovery of the REEs (Y,
Nd, Eu, Tb, Dy, and Er) was greater in mineral oil than with
kerosene as the organic solvent (Figure 3A). In contrast, the
SLM process demonstrated a consistently lower recovery using
mineral oil as a solvent (Figure 3B). This difference in recoveries
was approximately 18% for each REE in the case with 1% Span-
80 as the surfactant. For the SLM system, the opposite effect of
solvent was observed, with kerosene providing REE recovery
percentages that were an average 16% greater than REE recovery
percentages in the mineral oil SLM system. Previous work by
others43 with an amide chelating agent also has demonstrated
the impact of solvent on REE recovery. While mineral oil and
kerosene are both hydrocarbon mixtures, they may be
sufficiently different in polarity, resulting in observable variation
in REE mass transfer of the DEHP−metal complex.43 We also
note that we had significant difficulty in reproducing the mineral Figure 4. (A) Schematic of reactions during the separation and
concentration of REEs from coal fly ash leachate in the liquid emulsion
oil/DEHPA system for the SLM process. Despite using the same
membrane and the supported liquid membrane processes. (B)
experimental conditions, we sometimes observed zero recovery Concentration profile of a single REE (purple line) for a transport
of REEs (data not shown). The SLM separations with kerosene model where the steady state flux of the metal across each liquid (JL, JO,
were fairly reproducible (e.g., Figure 1C). While the reason for and JS) was controlled by diffusive transport across a diffusion film layer
the poor reproducibility of mineral oil in the SLM system of thickness δ at each liquid−liquid interface.
remains to be elucidated, we suspect that the miscibility of the
DEHPA in mineral oil might be a factor.
Model of REE Flux for Liquid Membrane Processes. strippant phases, the thickness of the oil layer for mass transfer,
The relative recovery of individual REEs differed greatly and the concentration of available chelating agent.
between the LEM and SLM processes. The lighter REEs tended In order to identify the potential influences of the system
configuration on overall recovery of different REEs, we
to be recovered more efficiently in the LEM process, while the
calculated the potential flux of each REE under steady state
heavier REEs were more efficiently extracted during the SLM
conditions as a means to understand the major rate-limiting step
process even though both processes used similar liquid−liquid for mass transfer. The model assumed that REE transport
reagents (10% DEHPA in organic solvent and 5 M nitric acid as occurred through five regimes: (i) bulk leachate, (ii) leachate
the strippant). The transfer of REE ions from the original ash film, (iii) organic phase, (iv) strippant film, and (v) bulk
leachate, into the oil phase and then into the acid stripping phase strippant as detailed in Figure 4B. The calculations were
entailed a multiple-step process (Figure 4A) including: (1) loss performed with the following assumptions:
of hydration shell surrounding the trivalent rare earth ion, (2)
1 Each flux (JL, JO, and JS) represents the flux of an
complexation of the rare earth ion with DEHPA, a process that individual metal.
depends on the relative binding constant between the metal ion 2 At steady state, the flux across the boundary layer of each
and the ligand, (3) mass transfer of the metal−DEHPA complex phase is equal (JL = JO = JS) and was equal to the product
through the oil phase, a process driven by diffusive flux across a of a diffusion coefficient and a concentration gradient
concentration gradient in the oil phase, (4) release of rare earth (Fick’s first law).
ion at the interface of the acid stripping solution and proton 3 The concentration of the fully protonated DEHPA in the
exchange with Me(DEHP·DEHPA)3, and (5) back diffusion of organic phase, (HR)2, is determined by equilibrium
protonated DEHPA to the oil-leachate interface. Any one of between the leachate and the organic phase.
these steps might be rate limiting for REE ion recovery in the 4 After extraction by DEHPA, all metals are then recovered
acid stripping step, depending on system configurations such as by the acid strippant with no metals remaining in the
interfacial surface areas between the leachate-oil and oil- organic phase.
F DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

5 The concentration of acid in the strippant phase remains (HR)2, the diffusion boundary layer thickness (δL, δO, and δS),
constant (5 M). and the diffusion coefficients DL, DO, and DS in the leachate, oil,
Complexation of all metals with DEHPA was assumed to and strippant phases. A complete summary of the methods used
occur according the following equilibrium reaction: to determine each individual parameter is given in the
Supporting Information.
Me3 + + 3(HR)2 ↔ Me(R·HR)3 + 3H+ (1′) Model calculations (Figure 5A) suggest that the flux of each
metal varies with the configuration of the system (i.e., LEM or
where the overbar indicates the organic solvent phase. DEHPA
dissolved in nonaqueous solvent exists as a dimer, (HR)2, and
the lanthanide−DEHPA complex exists as one lanthanide ion
with three DEHPA dimers, [Me(R·HR)3].44,45 For each
lanthanide, there is an associated equilibrium constant:
[Me(R· HR)3 ][H+]3
Keq = 3
[(HR)2 ] [Me3 +] (2′)
Equation 2 can be rearranged to create a relationship between
the metal concentrations at the interfacial regions of the organic
phase and both aqueous phases:
3
[Me(R· HR)3 ] Keq[(HR)2 ]
kL = =
[Me3 +]L [H+]L
3
(3a)
3
[Me(R· HR)3 ] Keq[(HR)2 ]
kS = =
[Me3 +]S [H+]S
3
(3b)
where subscripts L and S indicate the leachate and strippant
phases, respectively. The flux of an individual metal ion through
each phase is assumed to be driven by gradients in concentration
C as described by Fick’s First Law:
DL B Figure 5. Predicted initial fluxes for individual metal ions (normalized
JL = (C L − C LF) to the initial concentration in the leachate) for (A) both liquid
δL (4a)
membrane processes for a leachate with pH 4 (note log scale) and (B)
DO F the LEM process (same data points as part A) showing small variation
F
JO = (CO,L − CO,S) in flux due to differences in diffusion coefficients.
δO (4b)
DS F
JS = (CS − CSB) SLM), its ionic radius, and its binding affinity with DEHPA (i.e.,
δS (4c) Keq). The most important difference between the two systems is
the value for [(HR)2] within the organic phase. For the LEM
where the subscripts L, O, and S indicate the parameters for the
system at initial leachate pH 4, the calculated concentration of
leachate, organic, and strippant phases, respectively, and the
(HR)2 was 0.141 M, resulting in the following value for kL
superscripts B and F indicate the bulk and film layers,
(according to eq 3a′):
respectively. The diffusion coefficients D were calculated from
the hydrated ionic radius according to the Stokes−Einstein 3
[Me(R· HR)3 ] Keq[(HR)2 ]
relationship (Supporting Information). The diffusion film layer kL = =
δL and δS in the leachate and strippant phases were calculated [Me3 +]L [H+]L
3

from the mixing speed while δO was based on microscopic Keq(0.1410.141M)3


observations of oil film thickness in the emulsion or reported = = 2.81 × 109Keq
membrane thickness for the SLM system (Supporting (10−4)3 (3a′)
Information).
As a result of this (HR)2 concentration, at pH 4, the term kL is
By assuming that JL = JO = JS and using eqs 3a and 3b, an
approximately 109 or greater, depending on the value of Keq.
overall expression for the flux of an individual metal ion
normalized to its bulk concentration in the leachate can be Therefore, the kLδ L term in the denominator of eq 5 is
DL
expressed as δO k SδS
significantly greater than the and terms, resulting in the
DO DS
J kL
= following simplification:
C BL kLδ L
+
δO
+
k SδS
DL DO DS (5) kLδ L δ kδ J k D
for ≫ O and S S , L
= k δL = L
+ + DL DO DS CB δL
The variables kL and kS depend on [H ]L and [H ]S, respectively, L L
DL (6)
and both depend on the unchelated DEHPA concentration
[(HR)2]. Therefore, the normalized flux shown in eq 5 depends This simplification for the LEM system implies that normalized
on nine variables: the pH of the leachate and strippant phases fluxes for the REEs were limited by diffusion through the
(i.e., [H+]L and [H+]S), concentration of free chelating agent leachate film. The lanthanide contraction results in a decrease in
G DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

ionic radius with increase in atomic number along the lanthanide phase. For the SLM process, total volume requirements for the
series resulting in an increase in hydrated ionic radius and a organic solvent phase was orders of magnitude lower than
decrease of the diffusion coefficient. The diffusion coefficients conventional solvent extraction. Thus, REE recovery rate
along the lanthanide series were calculated to fall within an order depended mostly on relative affinity for the solvent carrier
of magnitude of each other and, therefore, yielded relatively (e.g., DEHPA). These findings are informative for future
small differences between fluxes among REEs (Figure 5B) that investigations of LEM or SLM processes that could focus on
was not sufficient to explain the observed results. A likely reason these components of the separation process through exper-
for this inconsistency is the model assumption of equal fluxes imentation with different chelating agents and the organic
across all three phases, an assumption that would not account for diluents.
observations of certain REE (e.g., Sc or Lu) immobilized on the Both LEM and SLM processes demonstrate high separation
organic phase and not recovered by the strippant phase as the factors for the REEs in the presence of other competing metal
equilibrium constant for Lu is 5 orders of magnitude larger than ion species. Future work should focus on further purification and
La. Thus in the LEM system, diffusion into the organic phase recovery of REEs from the strippant solutions in addition to
may be the overall rate-limiting step for the recovery of light reducing the total acid requirements. Full scale application of
REEs and dissociation of the Me(R·HR)3 complex may be rate- these processes will also require an understanding the effects of
limiting for the heavy REEs. pH for both leachate and strippant solution on REE recovery.
For the SLM system, the predicted concentration of (HR)2 at While this work was focused on leachates of coal fly ash, our
pH 4 was 9.9 × 10−6 M resulting in kL = 9.7 × 10−4 Keq. As a findings are applicable to other dilute mixed metal waste
consequence, the δO term was the largest term in the streams. Future work could focus on applying these methods to
DO other secondary feedstocks such as acid mine drainage,
denominator of eq 5 and flux simplified to the following: geothermal brines, medical/laboratory wastewater, or seawater.
These brine mixtures all pose similar challenges to coal ash
δO kδ kδ
for ≫ L L and S S , leachates in relation to matrix and difficulty in recovering REEs,
DO DL DS and liquid membrane processes could be advantageous due to
KeqDO[(HR)2 ]3 the relatively high selectivity for REE.


J kL kLDO
= = =
C BL δO
δO δO[H+]3 ASSOCIATED CONTENT
DO (7)
*
S Supporting Information
From this simplification, differences in flux between the REEs in The Supporting Information is available free of charge on the
the SLM system could be explained by variations in both the ACS Publications website at DOI: 10.1021/acs.est.9b00539.
metal-DEHPA binding constant, Keq, and the diffusion Schematic of the LEM and SLM processes, a table of the
coefficient DO. The value of DO decreases with increasing four coal ash sinters used as a basis for the synthetic
ionic radius (i.e., from light to heavy REE); however, the range leachate recipe, schematic of the SLM system, kinetics of
of values are within 1 order of magnitude. In contrast, the value SLM process, methods for correction of ICP-MS analysis,
of Keq increases by 5 orders of magnitude along the lanthanide method estimation of the solid phase composition of the
series. These results indicate the DEHPA concentration strippant after evaporation, effect of solvent choice on
[(HR)2] is smaller in the SLM system than the LEM, and that LEM process, and the methods used to estimate the
REE competition for (HR)2 is the key factor affecting recovery parameters for the flux model (PDF)
and flux in the SLM system. The limiting amount of free (HR)2 Spreadsheet of the initial and final concentrations of
is a direct result of the small total volume of solvent present in leachate and strippant for all processes (XLSX)


the membrane. As such, affinity toward free (HR)2 will be the
deciding factor for overall flux favoring heavier REEs over lighter
REEs. AUTHOR INFORMATION
While the flux calculations can compare the REEs for their Corresponding Authors
recovery potential, the model inadequately predicts observa- *E-mail: rs391@duke.edu.
tions of certain metals. For example, the model indicates a high *E-mail: hsukim@duke.edu.
flux of Sc. However, in our LEM and SLM separations, we ORCID
observed less than 1% of the original Sc in the strippant phase, Ryan C. Smith: 0000-0003-2591-416X
and a minimum of 40% Sc remaining in the original leachate James C. Hower: 0000-0003-4694-2776
depending on the method of recovery used. Sc was likely Mark R. Wiesner: 0000-0001-7152-7852
immobilized in the organic phase, probably because of the
Heileen Hsu-Kim: 0000-0003-0675-4308
strong binding affinity with DEHPA.42
Implications and Future Work. The unique chemical Notes
The authors declare no competing financial interest.


properties of REEs make them both critical and irreplaceable in
industrial applications. Due to the difficulty and environmental
impact of REE mining, developing methods for the recovery and ACKNOWLEDGMENTS
concentration of REEs from alternate sources such as coal ash is This work was supported by the U.S. National Science
crucial. Compared to the conventional kerosene-based solvent Foundation (No. CBET-1510965 and CBET-1510861) and
extraction, the LEM process was shown to be a potential the U.S. Department of Energy (DE-FE0026952).
alternative that could be implemented with mineral oil, a less
hazardous solvent than kerosene. Using the developed model,
we demonstrated that the recovery rates of REEs appeared to be
■ REFERENCES
(1) Van Gosen, B. S.; Verplanck, P. L.; Seal, R. R., II; Long, K. R.;
controlled by diffusional mass transfer of the REE across the oil Gambogi, J. Rare-earth elements. In Critical mineral resources of the

H DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

United StatesEconomic and environmental geology and prospects for (20) Diallo, M. S.; Kotte, M. R.; Cho, M. Mining Critical Metals and
future supply: U.S. Geological Survey Professional Paper 1802; Schulz, K. Elements from Seawater: Opportunities and Challenges. Environ. Sci.
J.; DeYoung, J. H., Jr.; Seal, R. R., II; Bradley, D. C., Eds.; U.S. Technol. 2015, 49, 9390−9399.
Geological Survey: Reston, VA, 2017; pp 01−031. (21) Taggart, R. K.; Hower, J. C.; Dwyer, G. S.; Hsu-Kim, H. Trends in
(2) Stegen, K. S. Heavy Rare Earths, Permanent Magnets, and the Rare Earth Element Content of U.S.-Based Coal Combustion Fly
Renewable Energies: An Imminent Crisis. Energy Policy 2015, 79, 1−8. Ashes. Environ. Sci. Technol. 2016, 50 (11), 5919−5926.
(3) Nassar, N. T.; Du, X.; Graedel, T. E. Criticality of the Rare Earth (22) American Coal Ash Association. 2016 Coal Combustion Product
Elements. J. Ind. Ecol. 2015, 19 (6), 1044−1054. (CCP) Production & Use Survey Report. https://www.acaa-usa.org/
(4) Riba, J. R.; López-Torres, C.; Romeral, L.; Garcia, A. Rare-Earth- publications/productionusereports.aspx (accessed October 9, 2018).
Free Propulsion Motors for Electric Vehicles: A Technology Review. (23) He, J.; Li, Y.; Xue, X.; Ru, H.; Huang, X.; Yang, H. Extraction of
Renewable Sustainable Energy Rev. 2016, 57, 367−379. Ce(IV) from Sulphuric Acid Solution by Emulsion Liquid Membrane
(5) Wang, Y.; Zhu, G.; Xin, S.; Wang, Q.; Li, Y.; Wu, Q.; Wang, C.; Using D2EHPA as Carrier. RSC Adv. 2015, 5 (91), 74961−74972.
Wang, X.; Ding, X.; Geng, W. Recent Development in Rare Earth (24) Lende, A. B.; Dinker, M. K.; Bhosale, V. K.; Kamble, S. P.;
Doped Phosphors for White Light Emitting Diodes. J. Rare Earths Meshram, P. D.; Kulkarni, P. S. Emulsion Ionic Liquid Membranes
2015, 33 (1), 1−12. (EILMs) for Removal of Pb(II) from Aqueous Solutions. RSC Adv.
(6) Alam, M. A.; Zuga, L.; Pecht, M. G. Economics of Rare Earth 2014, 4 (94), 52316−52323.
Elements in Ceramic Capacitors. Ceram. Int. 2012, 38 (8), 6091−6098. (25) Kandwal, P.; Mohapatra, P. K. A Novel Liquid Emulsion
(7) Mineral commodity summaries 2018; U.S. Geological Survey, U.S. Membrane Containing TODGA as the Carrier Extractant for Am
Government Printing Office: Washington, DC, 2017, DOI: 10.3133/ Recovery from Acidic Wastes. Sep. Sci. Technol. 2013, 48 (8), 1167−
70180197. 1176.
(8) Dutta, T.; Kim, K. H.; Uchimiya, M.; Kwon, E. E.; Jeon, B. H.; (26) Candela, A. M.; Benatti, V.; Palet, C. Pre-Concentration of
Deep, A.; Yun, S. T. Global Demand for Rare Earth Resources and Uranium (VI) Using Bulk Liquid and Supported Liquid Membrane
Strategies for Green Mining. Environ. Res. 2016, 150, 182−190. Systems Optimized Containing bis(2-Ethylhexyl) Phosphoric Acid as
(9) Binnemans, K.; Jones, P. T.; Blanpain, B.; Van Gerven, T.; Carrier in Low Concentrations. Sep. Purif. Technol. 2013, 120, 172−
Pontikes, Y. Towards Zero-Waste Valorisation of Rare-Earth- 179.
Containing Industrial Process Residues: A Critical Review. J. Cleaner (27) Kim, D.; Powell, L. E.; Delmau, L. H.; Peterson, E. S.;
Prod. 2015, 99, 17−38. Herchenroeder, J.; Bhave, R. R. Selective Extraction of Rare Earth
(10) Sprecher, B.; Xiao, Y.; Walton, A.; Speight, J.; Harris, R.; Kleijn, Elements from Permanent Magnet Scraps with Membrane Solvent
R.; Visser, G.; Kramer, G. J. Life Cycle Inventory of the Production of Extraction. Environ. Sci. Technol. 2015, 49 (16), 9452−9459.
Rare Earths and the Subsequent Production of NdFeB Rare Earth (28) Liang, P. E. I.; Liming, W. Study on a Novel Flat Renewal
Permanent Magnets. Environ. Sci. Technol. 2014, 48 (7), 3951−3958. Supported Liquid Membrane with D2EHPA and Hydrogen Nitrate for
(11) Jin, H.; Park, D. M.; Gupta, M.; Brewer, A. W.; Ho, L.; Singer, S. Neodymium Extraction. J. Rare Earths 2012, 30 (1), 63−68.
L.; Bourcier, W. L.; Woods, S.; Reed, D. W.; Lammers, L. N.; (29) Sengupta, B.; Sengupta, R.; Subrahmanyam, N. Process
Sutherland, J. W.; Jiao, Y. Techno-Economic Assessment for Intensification of Copper Extraction Using Emulsion Liquid Mem-
Integrating Biosorption into Rare Earth Recovery Process. ACS branes: Experimental Search for Optimal Conditions. Hydrometallurgy
Sustainable Chem. Eng. 2017, 5, 10148−10155. 2006, 84 (1−2), 43−53.
(12) Kose, B.; Cantoni, B.; Turolla, A.; Antonelli, M.; Hsu-Kim, H.; (30) Tunsu, C.; Petranikova, M.; Gergorić, M.; Ekberg, C.; Retegan,
Wiesner, M. R. Application of Nano Filtration for Rare Earth Elements T. Reclaiming Rare Earth Elements from End-of-Life Products: A
Recovery from Coal Fly Ash Leachate : Performance and Cost Review of the Perspectives for Urban Mining Using Hydrometallurgical
Evaluation. Chem. Eng. J. 2018, 349 (May), 309−317. Unit Operations. Hydrometallurgy 2015, 156, 239−258.
(13) Smith, Y. R.; Kumar, P.; McLennan, J. D. On the Extraction of (31) Smitha, V. S.; Surianarayanan, M.; Seshadri, H.; Lakshman, N. V.;
Rare Earth Elements from Geothermal Brines. Resources 2017, 6 (3), Mandal, A. B. Reactive Thermal Hazards of Tributyl Phosphate with
39−54. Nitric Acid. Ind. Eng. Chem. Res. 2012, 51 (21), 7205−7210.
(14) Ayora, C.; Macías, F.; Torres, E.; Lozano, A.; Carrero, S.; Nieto, (32) King, J. F.; Taggart, R. K.; Smith, R. C.; Hower, J. C.; Hsu-Kim,
J.-M.; Pérez-López, R.; Fernández-Martínez, A.; Castillo-Michel, H. H. International Journal of Coal Geology Aqueous Acid and Alkaline
Recovery of Rare Earth Elements and Yttrium from Passive- Extraction of Rare Earth Elements from Coal Combustion Ash. Int. J.
Remediation Systems of Acid Mine Drainage. Environ. Sci. Technol. Coal Geol. 2018, 195 (May), 75−83.
2016, 50, 8255−8262. (33) Taggart, R. K.; Rivera, N. A.; Levard, C.; Ambrosi, J.-P.;
(15) Stewart, B. W.; Capo, R. C.; Hedin, B. C.; Hedin, R. S. Rare Earth Borschneck, D.; Hower, J. C.; Hsu-Kim, H. Differences in bulk and
Element Resources in Coal Mine Drainage and Treatment Precipitates microscale yttrium speciation in coal combustion fly ash. Environ. Sci.
in the Appalachian Basin, USA. Int. J. Coal Geol. 2017, 169, 28−39. Processes Impacts 2018, 20, 1390−1403.
(16) Roth, E.; Bank, T.; Howard, B.; Granite, E. Rare Earth Elements (34) Bauer, D.; Diamond, D.; Li, J.; Sandalow, D.; Telleen, P.;
in Alberta Oil Sand Process Streams. Energy Fuels 2017, 31, 4714− Wanner, B. Critical Materials Strategy; U.S. Department of Energy:
4720. Washington, DC, 2010.
(17) Wilfong, W. C.; Kail, B. W.; Bank, T. L.; Howard, B. H.; Gray, M. (35) Ahmad, A. L.; Kusumastuti, A.; Derek, C. J. C.; Ooi, B. S.
L. Recovering Rare Earth Elements from Aqueous Solution with Porous Emulsion Liquid Membrane for Cadmium Removal: Studies on
Amine − Epoxy Networks. ACS Appl. Mater. Interfaces 2017, 9, 18283− Emulsion Diameter and Stability. Desalination 2012, 287, 30−34.
18294. (36) Fouad, E. A. Zinc and Copper Separation through an Emulsion
(18) Lin, R.; Stuckman, M.; Howard, B. H.; Bank, T. L.; Roth, E. A.; Liquid Membrane Containing Di-(2-Ethylhexyl) Phosphoric Acid as a
Macala, M. K.; Lopano, C.; Soong, Y.; Granite, E. J. Application of Carrier. Chem. Eng. Technol. 2008, 31 (3), 370−376.
Sequential Extraction and Hydrothermal Treatment for Character- (37) Chen, L.; Chen, J. Asymmetric Membrane Containing Ionic
ization and Enrichment of Rare Earth Elements from Coal Fly Ash. Fuel Liquid [A336][P507] for the Preconcentration and Separation of
2018, 232, 124−133. Heavy Rare Earth Lutetium. ACS Sustainable Chem. Eng. 2016, 4 (5),
(19) National Energy Technology Laboratory. Recovery of Rare Earth 2644−2650.
Elements from Coal and Coal Byproducts via a Closed Loop Leaching (38) Yan, X.; Dai, S.; Graham, I. T.; He, X.; Shan, K.; Liu, X.
Process : Final Report Recovery of Rare Earth Elements from Coal and Coal Determination of Eu Concentrations in Coal, Fly Ash and Sedimentary
Byproducts via a Closed Loop Leaching Process : Final Report, 7 June Rocks Using a Cation Exchange Resin and Inductively Coupled Plasma
2017. Mass Spectrometry (ICP-MS). Int. J. Coal Geol. 2018, 191, 152−156.

I DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

(39) Wu, S.; Zeng, X.; Dai, X.; Hu, Y.; Li, G.; Zheng, C. Accurate
Determination of Ultra-Trace Impurities, Including Europium, in
Ultra-Pure Barium Carbonate Materials through Inductively Coupled
Plasma − Tandem Mass Spectrometry. Spectrochim. Acta, Part B 2016,
123, 129−133.
(40) Robinson, P.; Townsend, A. T.; Yu, Z.; Mü n ker, C.
Determination of Scandium, Yttrium and Rare Earth Elements in
Rocks by High Resolution Inductively Coupled Plasma-Mass
Spectrometry. Geostand. Newsl. 1997, 23, 31−46.
(41) Tsurubou, S. Improved Extraction-Separation of Alkaline Earths
and Lanthanides Using Crown Ethers as Ion Size Selective Masking
Reagents : A Novel Macrocycle Application. Anal. Chem. 1995, 67,
1465−1469.
(42) Peppard, D. F.; Mason, G. W. Liquid-Liquid Extraction of the
Rare Earths and Related Elements. In Coordination Chemistry;
Kirschner, S., Ed.; Springer Science+Business Media, LLC.: New
York, 1969; pp 289−302.
(43) Cui, Y.; Yang, J.; Yang, G.; Xia, G.; Nie, Y.; Sun, G. Effect of
Diluents on Extraction Behavior of Rare Earth Elements with
N,N,N′,N′-Tetrabutyl-3-Oxy-Glutaramide from Hydrochloric Acid.
Hydrometallurgy 2012, 121−124, 16−21.
(44) Scharf, C.; Ditze, A.; Schwerdtfeger, K.; Kaufmann, D. E.;
Namyslo, J. C.; Fürmeier, S.; Bruhn, T. Investigation of the Structure of
Neodymium-Di-(2-Ethylhexyl) Phosphoric Acid Combinations Using
Electrospray Ionization and Matrix-Assisted Laser Desorption
Ionization Mass Spectrometry and Nuclear Magnetic Resonance
Spectroscopy. Metall. Mater. Trans. B 2005, 36 (4), 429−436.
(45) Kosinski, F. E. Lanthanum Solvent Extraction Using Di-(2-
Ethylhexyl) Phosphoric Acid. J. Inorg. Nucl. Chem. 1969, 31, 3623−
3631.

J DOI: 10.1021/acs.est.9b00539
Environ. Sci. Technol. XXXX, XXX, XXX−XXX

You might also like