Article 16 Rosita

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal Pre-proof

Sequential particle-size and magnetic separation for enrichment of


rare-earth elements and yttrium in Indonesia coal fly ash

Widya Rosita, I Made Bendiyasa, Indra Perdana, Ferian Anggara

PII: S2213-3437(19)30698-0
DOI: https://doi.org/10.1016/j.jece.2019.103575
Reference: JECE 103575

To appear in: Journal of Environmental Chemical Engineering

Received Date: 21 August 2019


Revised Date: 24 November 2019
Accepted Date: 29 November 2019

Please cite this article as: Rosita W, Made Bendiyasa I, Perdana I, Anggara F, Sequential
particle-size and magnetic separation for enrichment of rare-earth elements and yttrium in
Indonesia coal fly ash, Journal of Environmental Chemical Engineering (2019),
doi: https://doi.org/10.1016/j.jece.2019.103575

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2019 Published by Elsevier.


Sequential particle-size and magnetic separation for enrichment of rare-earth elements and
yttrium in Indonesia coal fly ash

Widya Rositaa,b*, I Made Bendiyasaa, Indra Perdanaa, Ferian Anggarac,d

a
Department of Chemical Engineering, Faculty of Engineering, Universitas Gadjah Mada, Jl Grafika no. 2,
Yogyakarta, 55281, Indonesia
b
Department of Nuclear Engineering and Engineering Physics, Faculty of Engineering, Universitas Gadjah Mada, Jl
Grafika no.2, Yogyakarta, 55281, Indonesia
c
Department of Geological Engineering, Faculty of Engineering, Universitas Gadjah Mada, Jl Grafika no. 2,
Yogyakarta, 55281, Indonesia

of
d
Unconventional Geo-resources Research Group, Faculty of Engineering, Universitas Gadjah Mada, Jl Grafika no.
2, Yogyakarta, 55281, Indonesia

ro
*
Corresponding author: widyar@ugm.ac.id

Abstract
Coal fly ash is known to contain rare-earth elements and yttrium (REY). In this research,

-p
three Indonesian coal fly ashes were characterized and subjected to sequential physical
separation processes. The characterization results showed that the ashes had a critical REY
content exceeding 38% and outlook coefficient (Coutlook) of 1.06–1.18; however, enrichment is
re
required prior to REY recovery. Coal fly ash from the Indramayu power plant also contained
high concentrations of Fe2O3 and CaO, which occurred as magnetite, hematite, spinel, and
srebrodolskite. Sequential physical separation consisting of sieving and magnetic separation was
lP

conducted to enrich the REY content of this ash. The sieving results showed that the REY
content increased with a decrease of particle size. The highest REY recovery (58.69%) and
enrichment factor (1.1) were obtained in ashes with a size fraction of < 38 µm. Magnetic
separation results showed that the nonmagnetic fraction contained more REY than the magnetic
na

fraction. The enrichment factor and REY recovery in the magnetic separation step were 1.11 and
91.64%, respectively. Overall REY recovery for the sequential separation process was 71.21%,
with an overall enrichment factor of 1.23. Sieving followed by magnetic separation effectively
decreased the Fe2O3 and CaO contents of this fly ash.
ur

Keywords:
Rare-earth Element, Coal fly ash, Particle size, Magnetic separation
Jo

1. Introduction
Rare-earth elements and yttrium (REY) are widely used and have become important
components in modern industries; however, although demand is high, availability of REY is very
limited [1–5]. Recycling waste materials that contain REY has become one alternative to
mitigate this deficiency [3, 6]. Researchers have found that coal fly ash contains a significant
amount of REY, so it has become of interest as a recycling source [5,7–9]. REY can be classified
into two groups based on their atomic number: light REY (LREY, which span the elements from
La to Sm) and heavy REY (HREY, from Eu to Lu) [3]. According to Seredin and Dai [10], REY
can be further categorized as critical (Nd, Eu, Tb, Dy, Y, and Er), uncritical (La, Pr, Sm, and

1
Gd), and excessive (Ce, Ho, Tm, Yb, and Lu). REY are non-volatile elements so their
distribution pattern in coal ash will be the same as that in coal. There are several distribution
patterns of REY enrichment in coal fly ash; namely, low (L), medium (M), and heavy (H) types.
These are compared with the upper crust continental (UCC) values, which are considered as N-
type. The distribution pattern is considered as L-type if LaN/LuN > 1, which originates from
terrigenous or tuffaceous sources, where REY enters the material at the peat-bog stage. The
pattern is M-type if LaN/SmN < 1 and GdN/LuN > 1, where the REY are supplied by hydrothermal
acid solution. H-type distribution occurs if LaN/LuN < 1, which arises due to the circulation of
natural water enriched in HREY in a coal basin [10].
In addition, Seredin and Dai [10] also state that the ideal coal ash for application as a
REY source should contain as many critical individual REY and as few excessive elements as
possible. The definitions of Outlook coefficient (Coutlook) and critical percentage are given by

of
Equations (1) and (2), respectively:

(𝑁𝑑 + 𝐸𝑢 + 𝑇𝑏 + 𝐷𝑦 + 𝐸𝑟 + 𝑌)/ ∑ 𝑅𝐸𝑌


𝐶𝑜𝑢𝑡𝑙𝑜𝑜𝑘 = (𝐶𝑒 + 𝐻𝑜 + 𝑇𝑚 + 𝑌𝑏 + 𝐿𝑢)/ ∑ 𝑅𝐸𝑌
; (1)

ro
(𝑁𝑑 + 𝐸𝑢 + 𝑇𝑏 + 𝐷𝑦 + 𝐸𝑟 + 𝑌)
𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 𝑝𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 = ∑ 𝑅𝐸𝑌
× 100. (2)

-p
In Indonesia, several coal-fired power plants produce fly ash. At present, the ashes are
mainly utilized as an admixture material in the cement industry and some are used for making
geopolymer [11]. The ashes are not currently treated to recover REY, although Anggara et al.
re
[12] showed that the REY content in Indonesian coal fly ash was tenfold higher than their
original content in the coal. Furthermore, Indonesian coal fly ash has a critical REY content
exceeding 38% and an H-type enrichment distribution pattern. Further investigation of recovery
lP

of contained REY from Indonesian coal fly ash is therefore needed.


REY recovery processes from coal fly ash include beneficiation and leaching.
Beneficiation processes include, among others, particle sizing, density separation, magnetic
separation, and froth flotation. Particle and magnetic separations have been used for coal fly ash
na

from USA, Poland, UK, and China [5, 9, 13–16]. Blisset et al. [5] found that REY from Polish
and UK coal fly ashes were more concentrated in the smaller size fractions of the nonmagnetic
material. This result was in agreement with those of Lin et al. [13, 14] and Dai et al. [9].
Different results were, however, reported by Zhang et al. [17], who studied three separation
ur

techniques—density separation, wet high-intensity magnetic separation, and froth flotation—for


fly ash and bottom ash samples collected from a utility site in the USA utilizing feed coal. The
results showed that the weak and moderately weak magnetic fractions contained a slightly higher
concentration of REY.
Jo

Lanzerstorfer [18] used air classification, instead of sieving, to produce particle size
fractions of Polish coal fly ash. Five size fractions were produced, with mass median diameters
of 2.2, 5.4, 9.7, 19.4, and 43.2 m. The Rare Earth Elements (REE) were enriched or depleted in
the different size fractions: all REYs were enriched by a factor of 1.05 to 1.65 in the finest
fraction, but depleted in the coarsest fraction by a factor of 0.63 to 0.78.
Literature from previous researchers shows some disagreements of experimental results,
which indicates that different coal fly ashes might give different responses to beneficiation. Coal
source is going to be a factor in fly ash quality, each fly ash is going to be unique. No studies of
the enrichment and recovery of REY from Indonesian coal fly ashes are reported. The purpose of

2
this study was therefore to evaluate the effect of particle-size and magnetic separations on REY
enrichment of such fly ashes.
2. Material and Methods
2.1. Material
Coal fly ashes were obtained from the discharges of the electrostatic precipitators of the
Tuban, Indramayu, and Paiton-1 coal-fired power plants in Indonesia, which are pulverized coal
combustion-type plants. The feedstocks were subbituminous coals from Kalimantan and South
Sumatera; detailed information concerning the mine locations was not available.

2.2. Physical separation


The physical separation process followed the method of dry sieving that was conducted
by Lin et al. [13]. Each sample was sieved using a sieve shaker. The following size ranges were
evaluated: > 75 m, 75–63 m, 63–53 m, 53–45 m, 45–38 m, < 38 m [13, 15, 19]. To

of
minimize agglomeration during dry sieving, only a small mass of sample (50–100 g) was
processed at a time. The sieving operation generally lasted for 75–90 min, until a constant
volume fraction was obtained.

ro
Once the optimal particle size fraction was obtained, wet magnetic separation was
performed using this fraction to increase the REY enrichment. Coal fly ash was first mixed with
water at a liquid-to-solid ratio of 4:1 (mL/g) and then placed into a wet magnetic separator.

-p
Magnetic flux densities of 4300 G and 10 800 G were applied. The magnetic component stuck to
the induced magnetic balls when the fly ash solution flowed around the balls in the magnetic
separator. The non-magnetic component exited at the bottom of the separator and then was
re
collected. After prescribed time, magnetic field around the ball was turned off. As a
consequence, magnetic component was released from the ball, flew to the bottom of separator
and then was collected. Magnetic and non-magnetic fractions were separated into their solid and
lP

liquid components. The solids were dried overnight in an oven at a temperature of 120 ºC. The
elemental composition and mineralogy of each size fraction, the magnetic and non-magnetic
fractions were analyzed.

2.3. Elemental, mineral and particle size analysis


na

Elemental contents of the fly ashes were analyzed using inductively coupled plasma mass
spectroscopy (ICP–MS) and atomic emission spectroscopy (ICP–AES). Major and trace
elements and loss on ignition (LOI) were assayed by ALS Geochemistry Laboratory, Kamloops,
British Columbia, Canada. For both major and trace element analysis, a prepared sample (0.1 g)
ur

was added to a lithium metaborate/lithium tetraborate flux, mixed well, and fused in a furnace at
1025 °C. The resulting melt was cooled and then dissolved in 100 mL of 4% nitric acid/2%
hydrochloric acid. This solution was analyzed by ICP–AES and ICP-MS and the results
Jo

corrected for spectral inter-element interferences. LOI was measured using a thermal
decomposition furnace or by thermogravimetric analysis. A prepared sample (1.0 g) was placed
in an oven at 1000 °C for 1 h, cooled, and then weighed. The percentage loss on ignition was
calculated from the difference in mass.
Mineralogy analysis of the fly ash samples was conducted by the GeoAssay–Mineralogy
Laboratory, PT Geoservices, Indonesia, using X-ray diffraction (XRD; D8 advance DaVinci;
Bruker, Germany), equipped with a cobalt X-ray tube operated at 30 kV and 50 mA. Qualitative
analysis was carried out by Rietveld refinement using Bruker Diffrac Suite Search/Match

3
software with the ICDD PDF-4 database. Scanning was carried out from 5º to 65º at increments
of 0.02°.
Petrographic analysis was performed in the Department of Geological Engineering,
Universitas Gadjah Mada, Indonesia, using a polished block sample of coal fly ash made using a
speciFix-20 Kit and Struers Labo System. Organic and inorganic component abundance
identifications were determined by using the genetic classification proposed by Hower [20].
Particle size analysis was carried out by the GeoAssay–Mineralogy Laboratory, PT
Geoservices, Indonesia using Laser Diffraction Particle Size Analyzer (LPSA; Malvern
Mastersizer3000).

3. Results and Discussion


3.1.Coal and coal fly ash characterization
Coal source information, combustion temperature, and total REY contents of the coals

of
corresponding to the fly ash samples taken from the three coal-fired power plants are presented
in Table 1. Their chemical compositions are displayed in Table 2.

ro
According to the ASTM C618 classification, the three fly ashes belong to Class F

-p
because the sum of the SiO2, Al2O3, and Fe2O3 contents exceeded 70%, while CaO was less than
10% [22]. Class F fly ash is often considered to be a product of higher-rank coal combustion,
while Class C ash results from combustion lower-rank coal [20]; however, the determining factor
is the major oxide composition, rather than the age or rank of the coal [20, 22, 23]. Although the
re
three fly ashes represented in this study were assigned to Class F according to their oxide
content, they originated from the combustion of lower-rank coals.
Table 3 summarizes the mineral compositions of the three coal fly ashes. They were
lP

dominated by quartz, mullite, and amorphous phases. This finding is supported by the data
displayed in Table 2, which shows high concentrations of SiO2 and Al2O3 compared with those
of other elements. The high Fe2O3 content could result from the Fe-bearing minerals (magnetite,
hematite, and spinel) in the fly ashes, which are generated from the decomposition and oxidation
na

of pyrite, siderite, and ankerite in the feed coals during the combustion process [24]. From a
thermodynamic perspective, the Gibbs free energy of the oxidation of magnetite to hematite is
favorable at lower temperatures, but not at higher temperatures. In addition, the equilibrium state
depends not only on temperature, but also the partial pressure of oxygen or a combination of
ur

these two variables. The oxidation of magnetite to hematite is favorable at higher partial
pressures of oxygen at a constant temperature [25]. This is in accordance with a report [26] that
mentions that magnetite starts to form at 827C, while the oxidation of magnetite to hematite is
slow. For this reason, both magnetite and hematite result from pyrite oxidation under rapid
Jo

heating conditions [26]. Table 3 shows that the amount of magnetite was higher than that of
hematite in all samples, although their combustion temperatures were different. The combustion
temperature in Paiton-1 (1100C) is lower than those of the Tuban and Indramayu power plants
(1400C and 1388C), but its fly ash magnetite content was higher than that of hematite. This is
most likely affected by the short combustion time of the particles when they pass the hot zone in
the boiler [27] or by the lower partial pressure of oxygen in the combustion process [25]. Other
possible reasons are the uneven temperature distribution within the particles or the formation of
hematite from oxidation of fine particles [28].

4
From the mineralogy data shown in Table 3, samples from Indramayu and Paiton-1 also
contained srebrodolskite (Ca2Fe3+2O5), which was not found in the Tuban sample. A new mineral
like srebrodolskite could have resulted from reactions between Fe from pyrite with Fe and Ca
from illite, carbonate, and clay, depending on the combustion time, temperature, and tendency to
react [29]. The high CaO content of the Indramayu and Paiton-1 samples (Table 2) would give
rise to a higher interaction probability between Fe and Ca with respect to the formation of
srebrodolskite.
Petrographic observation (Fig. 1) and observation of physical appearance showed that the
three samples had slightly different colors. The dark color was attributed to unburned carbon and
magnetite. Several studies show that there is considerable correlation between the color of fly
ash and its chemical and phase compositions [30, 31]. The ash color was mostly due to unburned
carbon and iron oxide. Unburned carbon is related to the gray and black colors of the fly ash,

of
while the characteristic color of iron oxides depends on the oxidation state of the iron. Magnetite,
which contains bivalent and trivalent iron, contributes the black color, but becomes brown when
evenly dispersed [31].

ro
-p
re
lP
na
ur
Jo

Fig. 1. Petrographic observation of original coal fly ash samples from (a) Tuban, (b) Indramayu,
and (c) Paiton-1 power plants

Table 4 shows the unburned carbon contents of the three fly ash samples. Unburned
carbon content is caused by incomplete combustion in the boiler [20]. Other literature states that

5
the unburned carbon content is also affected by the combustion system design; combustion
operating conditions; and characteristics of the coal feed, such as coal rank, the use of blended
coal, and the particle size [32]. The Tuban sample had the highest unburned carbon content.
Unburned carbon can inhibit the process of acid–base leaching in a subsequent REY recovery
process [32], because it will consume acid or base and change the pore structure and functional
groups of its surface [33]. REY recovery using acid–base leaching will need increased amounts
of the lixiviant. It is therefore necessary to reduce the amount of unburned carbon in coal ash
before its further processing to recover the REY content.

3.2. Sequential physical process for REY enrichment


The three coal fly ashes have critical REE values exceeding 38% and Coutlook values of
1.06–1.18, as shown in Table 2. The critical REY value found in this work was much higher
than that of Polish and United Kingdom coal fly ashes, which are in the range of 30%–36%, with

of
their Coutlook values within the range of 0.78–1.06 [5,7]. Despite this, the total REY contents of
these three coal fly ashes are still below the world average (445 ppm) [34], which indicates the
necessity of an enrichment treatment prior to REY recovery.

ro
To propose a suitable enrichment method, it is essential to know the mode of REY
occurrence in the fly ashes. Anggara et al. [35] investigated the mode of REY occurrence in
samples from Tuban and Indramayu using scanning electron microscopy with energy-dispersive

-p
spectroscopy (SEM–EDS), the results of which showed that the REY were found in an
aluminosilicate glass phase and spinel, but were not present in the unburned carbon. REY
occurrence in aluminosilicate glass was also found by Hower et al. [8] using wavelength-
re
dispersive spectrometry electron microprobe analysis of epoxy-bound polished pellets, by
Thompson et al. [36] using laser ablation (LA)–ICP–MS, and by Kolker et al. [37] using
SHRIMP–RG (sensitive high-resolution ion microprobe–reverse geometry). Recent studies show
lP

that nanoscale examination is important to understand the associations of REE in fly ashes.
Transmission electron microscopy (TEM) analysis of Central Appalachian coal-derived fly ashes
indicated that REY were found in zircon, monazite, and within the glass. Subsequent high-
resolution (HR)–TEM analysis of Al–Si glass showed that REE were associated with the glass
and were also found in particles within the amorphous carbon surrounding the glass [38]. Other
na

work using SEM, TEM, and selected-area electron diffraction (SAED) on fly ash from the
combustion of eastern Kentucky Fire Clay coal showed that REE were distributed within glassy
particles and, in certain cases, were accompanied by phosphorus [39]. To better understand the
detailed mode of occurrence of REY in coal fly ash, a sequential chemical extraction procedure
ur

and the use of SEM–EDS followed by establishment of a model for REY–major element
correlation using the Statistical Package for Social Science (SPSS) software was also
investigated [15, 16]. The results of this study showed, through Pearson correlation and
Jo

regression correlation analysis, that the REY content was strongly and positively correlated with
the silicon and aluminum contents, but negatively correlated with Fe2O3 and S. The dominant
mode of occurrence of REY was in the silicate and aluminosilicate forms [15, 16]. Cao et al. [40]
also reported the results of REY leaching from coal fly ash, which was assumed to be in the form
of spherical particles, using HCl as the lixiviant. A sequential phenomenon was observed in this
reaction: REY first dissolved from the particle surface of the fly ash powder, followed by REY
entrapped in the vitreous body. Kinetic analysis of the leaching showed that La was more easily
leached than Ce and Nd. The leachability for REY elements is mainly due to the association of
each REY with different minerals, as well as their ionic radius differences [40].

6
Table 2 shows that the Indramayu sample had the highest critical REY value and Coutlook
of the three fly ash samples. It also had higher Fe2O3 content than those from Tuban and Paiton-
1. The sample from Indramayu coal-fired power plant was therefore chosen to proceed with
REY-enrichment processing using sequential physical separation. Unburned carbon was first
reduced by sieving and enrichment was continued by magnetic separation. It was reported that
sieving does not lead to sample contamination and can be used for separating unburned carbon
from fly ash [32]. According to Thompson et al. [36], magnetic separation might be an effective
way of concentrating non-magnetic REE phases. Major element concentration changes were also
observed after sieving [9,18] and magnetic separation process [41,42].
REY was associated with pure aluminosilicate glass, Ca-Fe enriched aluminosilicate
glass and several other minerals that contain major elements [8,15,16,36,37,40]. This suggests
that major element concentration has an influence in REY enrichment process. Information about
the presence of major elements in particular fly ash is necessary to further enrich the REY

of
content.
As reported by Lin et al. [13], REY enrichment could be quantitatively characterized
using the enrichment factor (EF), while efficiency of the separation method was quantified using

ro
REY recovery. For the sieving process, EF was characterized using Equation (3) and the REY
efficiency using Equation (4). Equations (5) and (6) were similarly applied to evaluate the
magnetic separation process. The overall EF and REY recovery were evaluated using Equations

-p
(7) and (8), respectively. Equations (5) to (8) are modifications of those developed by Lin et al.
[13].
re
𝑅𝐸𝑌
𝐸𝐹𝑖 = 𝑅𝐸𝑌 𝑖 ; (3)
𝑓
(𝑅𝐸𝑌𝑖 𝑊𝑖 )
𝑅𝑒𝑐𝑜𝑣𝑒𝑟𝑦𝑖 (%) = × 100; (4)
(∑𝑘
𝑖 𝑅𝐸𝑌𝑖 𝑊𝑖 )
lP

𝑅𝐸𝑌𝑛𝑚
𝐸𝐹𝑛𝑚 = ; (5)
𝑅𝐸𝑌𝑓𝑛
(𝑅𝐸𝑌𝑛𝑚 𝑊𝑛𝑚 )
𝑅𝑒𝑐𝑜𝑣𝑒𝑟𝑦𝑛𝑚 (%) = × 100; (6)
(𝑅𝐸𝑌𝑓𝑛 𝑊𝑓𝑛 )
𝑅𝐸𝑌𝑛𝑚
𝐸𝐹𝑜 = ; (7)
na

𝑅𝐸𝑌𝑓
(𝑅𝐸𝑌𝑛𝑚 𝑊𝑛𝑚 )
𝑅𝑒𝑐𝑜𝑣𝑒𝑟𝑦𝑜 (%) = × 100. (8)
(∑𝑘
𝑖 𝑅𝐸𝑌𝑖 𝑊𝑖 )

REYi is the total REY concentration (ppm), where the subscript i represents the ith size fraction;
ur

REYf is the total REY concentration (ppm) in the feed to the sieving process; Wi is the mass of
the ith size fraction (kg); REYiWi is the mass of total REY in the ith size fraction (mg); k is the
total number of fractions. REYnm is the total REY concentration (ppm) in nonmagnetic (nm)
Jo

fraction; REYfn is the total REY concentration (ppm) in the feed to the magnetic separation
process; Wfn is the mass of the feed to the magnetic separation process (kg); Wnm is the mass of
the nonmagnetic fraction (kg); REYnmWnm and REYfnWfn are the masses of total REY in the
nonmagnetic fraction and in the feed to the magnetic separation process (mg), respectively.
The particle size separation and magnetic separation are discussed separately in the
following sections.

3.2.1. Particle size separation

7
Fig. 2 shows that Indramayu fly ash was dominated by the fine fraction of < 38 m. This
result is similar to those for fly ashes from Ohio [13] and Kentucky [41] pulverized-coal-fired
power plants and a Polish coal-fired power plant [18].

of
ro
-p
re
Fig. 2. Particle size distribution in Indramayu sample of coal fly ash, as determined by Laser
Diffraction Particle Size Analyzer (LPSA).
lP

In this study, the coal fly ash was separated using dry sieving method. In order to
determined separation efficiency using this technique, each fraction was being collected and each
volume fraction was measured. The volume distribution is then compared with the data resulted
from LPSA analysis, as shown in Table 5.
na

Table 5 shows that results from LPSA and dry sieving indicates a similar trend in particle
size distribution. However, between the two measurements there is a significant different amount
of retained volume in the 53-45 m and 45-38 m fraction. The difference might be due to
ur

agglomeration that starts to take place for the finer particles. The finer particles seem to become
more hygroscopic and therefore easier to stick to each other. As a result, in the dry sieving some
amounts of fine particles are agglomerated and they then are retained in 53-45 m and 45-38 m
fractions. Meanwhile, the larger portion of finest particle can still pass through the sieving size of
Jo

< 38 m. This result suggests that dry sieving method can still be applied for separating particle
size < 38 m.

8
of
ro
-p
re
Fig. 3. Composition and loss on ignition (LOI) of major elements (reported as oxides) in
Indramayu sample of coal fly ash.
lP

Fig. 3 shows that particles in the > 75 μm size fraction had the highest LOI content. LOI
can be used as a simple indicator to detect the existence of unburned carbon, although there are
many factors that can affect this value. Consequently, other methods are required to measure the
na

amount of unburned carbon in the samples [22, 27].


Fig. 4 shows that < 38 m particle size fraction also had the lowest amorphous phase
content. This originated from aluminosilicate glass and unburned carbon. XRD analysis could
not, however, quantitatively differentiate the contribution of unburned carbon to the total
amorphous phase content, so petrographic observation was used to evaluate the > 75 m, 53–45
ur

m, and < 38 m particle size fractions. From the petrographic analysis shown in Fig. 5, it was
found that the unburned carbon in each fraction contributed as much as 22%, 14%, and 7%,
respectively, of the amorphous phases.
Jo

9
of
ro
-p
re
Fig. 4. Mineral compositions of various particle size fractions in Indramayu sample of coal fly
ash, as determined by X-ray diffraction.
lP
na
ur
Jo

Fig. 5. Petrographic composition of various particle size fractions in Indramayu coal fly ash
sample.

The unburned carbon content in each fraction was compared with that in the original coal
fly ash (Table 4). There were increases of unburned carbon content in the > 75 m and 53–45

10
m particle size fractions from 13% to 22% and 14%, respectively, but a depletion of unburned
carbon content (7%) in the < 38 m size fraction. The results indicate that unburned carbon
could be reduced using sieving and collected in the coarser size particles. Fig. 6 shows that a
coarser size of unburned carbon was observed in the > 75 m particle size, while the finer size
was dispersed in particles of < 38 m. For more effective removal of unburned carbon, other
techniques should be applied, such as froth flotation [42, 43] or thermal processes [44].

of
ro
-p
re
Fig. 6. Petrographic compositions of various particle size fractions: (a) > 75 m; (b) 53–45 m;
lP

(c) < 38 m.

Table 6 presents the chemical compositions of various particle size ranges. The coarser
particle sizes had higher SiO2 and lower CaO contents, whereas those of Al2O3 and Fe2O3
na

remained relatively constant. Fig. 4 shows that higher mullite and quartz contents were present in
the coarser fraction, but this contained less spinel, magnetite, hematite, and srebrodolskite. Fig. 5
shows that the coarser size particles also contained more unburned carbon than the
aluminosilicate glass. The high SiO2 content in the coarser size particles (Table 6) was assumed
ur

to originate mostly from quartz and mullite (3Al2O3.2SiO2), with some contribution from the
aluminosilicate glass (Figs. 4 and 5). The Al2O3 content, which was relatively constant
throughout all particle size fractions, was mostly contributed by mullite, spinel, and
Jo

aluminosilicate glass. Although the magnetite, hematite, and srebrodolskite phases increased
with decreasing particle size fraction, the Fe2O3 content remained constant (Table 6). This
indicated that iron was also present in the aluminosilicate glass. This is probably because glass
and Fe-bearing minerals can occur in the same particle, as reported by other researchers [37, 38,
44]. Table 6 shows that the CaO content increased in the smaller particle sizes, which resulted
from the increase in the carbonate and srebrodolskite contents in the fine particles, as depicted in
Fig. 4. A significant increase in the finest particle size fraction may have originated from Ca,
which was incorporated in aluminosilicate glass. Ca–Fe-bearing aluminosilicate glass
(aluminosilicate glass matrix) was also found in ash in a study conducted by Kolker et al. [37].

11
The REY content was highest in the finest particle size fraction, as presented in Table 6.
In the present work, the smallest particle size (< 38 m fraction) contained a greater proportion
of aluminosilicate glass matrix. Previous research conducted on the Indramayu sample by
Anggara et al. [35] found that most REY occurred in the glass. As a result, if glass is abundant,
the REY content will be also high in this fraction. This result is in agreement with those of
previous reports [5, 13, 15, 16, 41, 45].

of
ro
-p
re
lP

Fig. 7. Enrichment factor and REY recoveries of various particle size fractions of Indramayu
coal fly ash sample.

The effect of particle size on EF and REY recovery are shown in Fig. 7. Both parameters
generally increased as the particle size decreased. This result is in accordance with those
na

obtained by Dai et al. [9] and Lanzerstorfer et al. [47]. The REY recovery significantly increased
for the < 38 m size fraction to 58.69%, while the EF linearly increased with a decrease of
particle size.
Based on the above discussion, the < 38 m particle size fraction had the highest REY
ur

content, so this was chosen for the subsequent enrichment process; however, it also contained
significant amounts of iron oxide, which diluted the REY concentrations. Iron-bearing minerals
are magnetic, so removal of these minerals by magnetic separation would further increase the
Jo

REY concentrations. Another beneficial effect of removing iron-bearing minerals is the increase
in REY recovery in the subsequent leaching process, because an acidic lixiviant is generally
used. Several studies have shown that iron oxides can react with acidic leaching agents [47, 48],
so its removal is expected to increase the efficiency of REY recovery by leaching.

3.2.2 Magnetic separation process


Based on the discussion in Section 3.2.1, REY in the nonmagnetic fraction can be
enriched using a magnetic separation; however, because finer coal fly ash is hygroscopic, the
particles are easy agglomerated, so the nonmagnetic fraction is likely to attach to the magnetic

12
fraction. Wet magnetic separation is therefore appropriate to separate the nonmagnetic and
magnetic fractions [21].
Magnetic flux densities of 4300 G and 10 800 G were used. The effect of this parameter
on the composition of the major compounds in the < 38 m size fraction of the Indramayu
sample is shown in Fig. 8. The SiO2 and Al2O3 contents increased in the nonmagnetic fraction
resulting from magnetic separation using these flux densities. SiO2 mostly originated from quartz
and mullite; Al2O3 was derived from mullite and spinel; the aluminosilicate glass matrix
contributed both SiO2 and Al2O3. In contrast, the nonmagnetic fraction contained less Fe2O3,
CaO, and MgO. The Fe2O3 was mostly derived from Fe-bearing minerals, with some from the
aluminosilicate glass matrix. This indirectly indicated that the magnetic fraction contained more
Fe-bearing minerals than the nonmagnetic fraction. Similar trends have been reported by other
researchers [43, 49].

of
ro
-p
re
lP
na
ur
Jo

Fig. 8. Major oxide contents in nonmagnetic fraction after separation at two magnetic flux
densities.
Fig. 8 also shows that the non-magnetic fraction resulting from separation at 10 800 G
contained higher SiO2, constant Al2O3, and lower Fe2O3 and CaO contents than that produced at
4300 G. The SiO2, Al2O3, Fe2O3, and CaO contents of the nonmagnetic products at 10 800 and
4300 G were 47.8%, 23.6%, 8.4%, 7.8%, and 44.6%, 23.5%, 10.2%, 8.9%, respectively. The
lower Fe2O3 content in the nonmagnetic fraction indicates that Fe-bearing minerals were more

13
strongly attracted to the magnetic fraction. According to a study of nineteen US and international
coal fly ash samples conducted by Kolker et al. [37] using a SHRIMP–RG ion microprobe, two
kinds of aluminosilicate glass occurred: one consisted almost entirely of Al and Si (pure
aluminosilicate glass); the other was a Ca–Fe-enriched aluminosilicate glass. In this study,
separation using 10 800 G was predicted to attract more Ca–Fe-enriched aluminosilicate glass to
its magnetic fraction; consequently, the nonmagnetic products contained more amount of pure
aluminosilicate glass. The presence of this pure aluminosilicate glass, quartz, and mullite led to
increased SiO2 and Al2O3 contents in the nonmagnetic fraction for separation at 10 800 G.
The REY recovery to the nonmagnetic fraction using a magnetic flux density of 10 800 G
was lower (71.63%) than that using 4300 G (91.64%), as shown in Table 7. The non-magnetic
fraction resulting from separation at 10 800 G contained more amount of pure aluminosilicate
glass, quartz, and mullite. According to Kolker et al. [37], REY are partitioned between pure and
Ca–Fe-enriched aluminosilicate glasses. Fe-bearing minerals were also found to be REY-

of
bearing, while quartz tended to be REY-depleted. Because Ca–Fe-enriched aluminosilicate glass
and Fe-bearing minerals were more strongly attracted to the magnetic fraction after separation at
10 800 G than at 4300 G, REY residing in these phases also reported to the magnetic fraction.

ro
The total REY in the nonmagnetic fraction was consequently lower for separation at 10 800 G.
This explains the lower REY magnetic separation recovery at higher magnetic flux density.

-p
The chemical composition and mineralogy of products from the 4300 G separation
process are discussed in this section. Inspection of the chemical compositions of the magnetic
and nonmagnetic fractions shows that SiO2 and Al2O3 contents were higher in the latter, as
re
shown in Fig. 9. SiO2 and Al2O3 in the nonmagnetic fractions mostly appeared as quartz and
mullite phases, as shown in Fig. 10. Fig. 9 shows that the CaO, Fe2O3, and MgO contents were
lower in the non-magnetic fraction. These elements were represented by magnetite, hematite,
lP

srebrodolskite, and spinel phase, as illustrated in Fig. 10. As previously discussed, SiO2, Al2O3,
CaO, Fe2O3, and MgO also originated from the aluminosilicate glass matrix. An pure Al–Si glass
phase will contribute SiO2 and Al2O3 content, while a Ca–Fe-enriched aluminosilicate glass will
affect CaO, Fe2O3, and MgO contents. Unfortunately, the data shown in Fig. 10 and Table 8
cannot differentiate between these two aluminosilicate glasses, so further experiments were
na

required. Fig. 9 shows that CaO content in the non-magnetic fraction was 8.9%, which is
considered typical of a low-calcium coal fly ash. It is therefore suggested that an alkaline leach
followed by an acid leach would be appropriate for further REY recovery by leaching [51].
ur
Jo

14
of
ro
-p
Fig. 9. Contents of major oxides in nonmagnetic and magnetic products after magnetic
re
separation using 4300 G.
lP
na
ur
Jo

Fig. 10. Mineral composition of nonmagnetic products after magnetic separation at 4300 G, as
determined by X-ray diffraction.

The effect of magnetic separation on REY enrichment is shown in Table 7. According to


Wills et al. [52] both EF and recovery are used for evaluating the concentrating process of REY,
even though both of them are independent of each other. There is an inverse relationship between

15
EF and recovery. If the high EF value is desirable, the process will result in more tailing and
recovery will be low. On the contrary, high recovery of REY can be obtained, but its concentrate
still contains more impurities, so the EF value will decrease. Process that having greater values
of EF and recovery than the others is selected.
The EF and REY recovery were higher in the non-magnetic fraction, where the REY
content was probably mostly incorporated with the pure aluminosilicate glass phase. The REY
mostly accumulated in the Ca–Fe-enriched aluminosilicate glass in the magnetic phase. The EF
obtained by magnetic separation was 1.12 and the REY recovery was 91.64% (Table 7). This
result is in line with that of Lin et al. [13], who reported EF values of 1.01–1.13 and recoveries
of 49%–96%. To evaluate the overall performance of the two-step sequential physical separation
processes on REY enrichment, the overall EF and REY recovery were calculated using
Equations (5) and (6), respectively. These parameters were 1.23 and 71.21%, respectively. If EF
and recovery values resulted from magnetic separation and from sequential separation were

of
compared, sequential separation shows higher value of EF and lower value of recovery. The
increasing of EF value is caused by gradual concentrating process. However, sequential
separation resulted in more tailing than that from magnetic separation process only.

ro
The value of overall EF and recovery (1.23 and 71.21%, respectively) is implied that a
subsequent chemical process is required to increase the REY content in fly ash.

-p
Conclusions
Three Indonesian coal fly ashes were characterized as potential REY sources. An REY
enrichment process is required before they can be recovered. Of the three samples, that
re
originating from South Sumatera coal combustion in the Indramayu coal-fired power plant had a
unique character because of its relatively high contents of Fe2O3 and CaO. It was found that the
REY content increased as the particle size of the ash decreased. The highest EF and REY
recovery of 1.1 and 58.69%, respectively, resulted from the finest size fraction (< 38 μm). The
lP

Fe2O3 content was also enriched in the finest fraction, so magnetic separation was used to
separate the REY-enriched fraction. An EF of 1.12 and REY recovery of 91.64% were obtained
for the magnetic separation process. By sequential sieving and magnetic separation, the overall
EF reached a value of 1.23, while the overall REY recovery was 71.21%.
na

Author contributions :
ur

Widya Rosita was responsible for the collection of the original coal fly ash sample; Widya Rosita, I Made
Bendiyasa, Indra Perdana and Ferian Anggara were responsible for designing the sequential physical
Jo

separation process; Widya Rosita and Indra Perdana were responsible for the collection of particle size
separation sample; Widya Rosita and I Made Bendiyasa were responsible for the collection of magnetic
separation sample; Ferian Anggara conducted the petrographic examination of the sample; all authors
participated in the writing, review, and editing of the manuscript.

16
Acknowledgment

This work was financially supported by the Ministry of Research, Technology and Higher
Education, Republic of Indonesia. The authors would like to thank Ajrun Karim, Elsandy Adha
Mukhti, Isyatun Rodliyah, M Ikhsan Kurniawan, Dea Anisa Ayu Besari, and Hotden Manurung
for their support in this research. We are also grateful to Dr. Himawan Tri Bayu M.P, Dr. Agus
Prasetyo, Dr. Sutijan, Dr. Maykel T.E Manawan, Dr. Alexander Agung for insightful discussion.
We also thanks to anonymous reviewers for their valuable comments in order to improve this
manuscript.

of
References

ro
[1] E. Alonso, A.M. Sherman, T.J. Wallington, M.P. Everson, F.R. Field, R. Roth, R.E.
Kirchain, Evaluating Rare Earth Element Availability: A Case with Revolutionary
Demand from Clean Technologies, Environ. Sci. Technol. 46 (2012) 3406–3414.

-p
doi:10.1021/es203518d.
[2] A. Golev, M. Scott, P.D. Erskine, S.H. Ali, G.R. Ballantyne, Rare Earths Supply Chains:
Current Status, Constraints and Opportunities, Resour. Policy. 41 (2014) 52–59.
re
doi:10.1016/j.resourpol.2014.03.004.
[3] M. Simoni, E.P. Kuhn, L.S. Morf, R. Kuendig, F. Adam, Urban Mining as a Contribution
to the Resource Strategy of the Canton of Zurich, Waste Manag. 45 (2015) 10–21.
doi:10.1016/j.wasman.2015.06.045.
lP

[4] J.W. Ahn, T. Thriveni, Y. Jegal, Occurrence and Distribution of Rare Earths with
Different Coal Power Plants Ash and Recovery of Critical Rare Earths from Coal Ash for
Simultaneous Utilization of CO2, in: World of Coal Ash (WOCA) Conference, University
of Kentucky Center for Applied Energy Research, Nashville, TN, 2015.
na

http://www.flyash.info/2015/038-Ahn-2015.pdf.
[5] R. Blissett, N. Smalley, N.A. Rowson, An Investigation into Six Coal Fly Ashes from the
United Kingdom and Poland to evaluate rare earth element content, Fuel. 119 (2014) 236–
239. doi:10.1016/j.fuel.2013.11.053.
ur

[6] K. Binnemans, P.T. Jones, B. Blanpain, T. Van Gerven, Y. Yang, A. Walton, M. Buchert,
Recycling of Rare Earths: A critical Review, J. Clean. Prod. 51 (2013) 1–22.
doi:10.1016/j.jclepro.2012.12.037.
Jo

[7] W. Franus, M.M. Wiatros-Motyka, M. Wdowin, Coal Fly Ash as a Resource for Rare
Earth Elements, Environ. Sci. Pollut. Res. 22 (2015) 9464–9474. doi:10.1007/s11356-015-
4111-9.
[8] J.C. Hower, J.G. Groppo, P. Joshi, S. Dai, D.P. Moecher, M.N. Johnston, Location of
Cerium in Coal-Combustion Fly Ashes : Implications for Recovery of Lanthanides, Coal
Combustion Gasification Products. 5 (2013) 73–78. doi:10.4177/CCGP-D-13-00007.1.
[9] S. Dai, L. Zhao, S. Peng, C.L. Chou, X. Wang, Y. Zhang, D. Li, Y. Sun, Abundances and
Distribution of Minerals and Elements in High-Alumina Coal Fly Ash from The Jungar
Power Plant, Inner Mongolia, China, Int. J. Coal Geol. 81 (2010) 320–332.
doi:10.1016/j.coal.2009.03.005.

17
[10] V. V. Seredin, S. Dai, Coal Deposits as Potential Alternative Sources for Lanthanides and
Yttrium, Int. J. Coal Geol. 94 (2012) 67–93. doi:10.1016/j.coal.2011.11.001.
[11] R. Fatikhin, R.B. Cahyono, H.T. Petrus, I. Perdana, Synthesis of dry-mix of fly ash based
geopolymer, in: T. Ariyanto, Rochmadi, I. Prasetyo, N.R. Eviana (Eds.), AIP Conf. Proc.,
AIP Publishing, Yogyakarta, 2019: p. 020050. doi:https://doi.org/10.1063/1.5095028.
[12] F. Anggara, D.H. Amijaya, A. Harijoko, T.N. Tambaria, A.A. Sahri, Z.A.N. Asa, Rare
Earth Element and Yttrium Content of Coal in The Banko Coalfield, South Sumatra
Basin, Indonesia: Contributions from Tonstein Layers, Int. J. Coal Geol. 196 (2018) 159–
172. doi:10.1016/j.coal.2018.07.006.
[13] R. Lin, B.H. Howard, E.A. Roth, T.L. Bank, E.J. Granite, Y. Soong, Enrichment of Rare
Earth Elements from Coal and Coal by-Products by Physical Separations, Fuel. 200
(2017) 506–520. doi:10.1016/j.fuel.2017.03.096.
[14] R. Lin, M. Stuckman, B.H. Howard, T.L. Bank, E.A. Roth, M.K. Macala, C. Lopano, Y.

of
Soong, E.J. Granite, Application of sequential extraction and hydrothermal treatment for
characterization and enrichment of rare earth elements from coal fly ash, Fuel. 232 (2018)
124–133. doi:10.1016/j.fuel.2018.05.141.

ro
[15] J. Pan, C. Zhou, M. Tang, S. Cao, C. Liu, N. Zhang, M. Wen, Y. Luo, T. Hu, W. Ji, Study
on the modes of occurrence of rare earth elements in coal fly ash by statistics and a
sequential chemical extraction procedure, Fuel. 237 (2019) 555–565.

-p
doi:10.1016/j.fuel.2018.09.139.
[16] J. Pan, C. Zhou, C. Liu, M. Tang, S. Cao, T. Hu, W. Ji, Y. Luo, M. Wen, N. Zhang,
Modes of Occurrence of Rare Earth Elements in Coal Fly Ash: A Case Study, Energy
re
Fuels. 32 (2018) 9738–9743. doi:10.1021/acs.energyfuels.8b02052.
[17] W. Zhang, J. Groppo, R. Honaker, Ash Beneficiation for REE Recovery, in: World of
Coal Ash (WOCA) Conference, University of Kentucky Center for Applied Energy
lP

Research, Nashville, TN, 2015. http://www.flyash.info/194-Groppo-2015.pdf.


[18] C. Lanzerstorfer, Fly ash from coal combustion : Dependence of the concentration of
various elements on the particle size, Fuel. 228 (2018) 263–271.
[19] E. Roth, R. Lin, B.H. Howard, T.L. Bank, E.J. Granite, Y. Soong, Distributions and
Extraction of Rare Earth Elements from Coal and Coal By-Products, in: World of Coal
na

Ash (WOCA) Conference, University of Kentucky Center for Applied Energy Research,
Lexington,KY, 2017: p. 103. http://www.flyash.info/2017/113-Roth-woca2017p.pdf.
[20] J.C. Hower, Petrographic Examination of Coal-Combustion Fly Ash, Int. J. Coal Geol. 92
(2012) 90–97. doi:10.1016/j.coal.2011.12.012.
ur

[21] N.D. Nugroho, W. Rosita, I. Perdana, I.M. Bendiyasa, F.R. Mufakhir, W. Astuti, Iron
bearing oxide minerals separation from rare earth elements (REE) rich coal fly ash, in:
IOP Conf. Ser. Mater. Sci. Eng., IOP Publishing, 2019. doi:10.1088/1757-
Jo

899X/478/1/012026.
[22] Z. Wang, S. Dai, J. Zou, D. French, I.T. Graham, Rare earth elements and yttrium in coal
ash from the Luzhou power plant in Sichuan, Southwest China: Concentration,
characterization and optimized extraction, Int. J. Coal Geol. 203 (2019) 1–14.
doi:10.1016/J.COAL.2019.01.001.
[23] S. Dai, V. V. Seredin, C.R. Ward, J. Jiang, J.C. Hower, X. Song, Y. Jiang, X. Wang, T.
Gornostaeva, X. Li, H. Liu, L. Zhao, C. Zhao, Composition and modes of occurrence of
minerals and elements in coal combustion products derived from high-Ge coals, Int. J.
Coal Geol. 121 (2014) 79–97. doi:10.1016/j.coal.2013.11.004.

18
[24] S. V. Vassilev, C.G. Vassileva, A.I. Karayigit, Y. Bulut, A. Alastuey, X. Querol, Phase-
Mineral and Chemical Composition of Fractions Separated from Composite Fly Ashes at
the Soma Power Station, Turkey, Int. J. Coal Geol. 61 (2005) 65–85.
doi:10.1016/j.coal.2004.05.004.
[25] D.R. Gaskell, Metal Production: Ellingham Diagrams, in: Encycl. Mater. Sci. Technol.,
Elsevier, 2001: pp. 5481–5486. doi:10.1016/B0-08-043152-6/00956-6.
[26] J. Tomeczek, H. Palugniok, Kinetics of mineral matter transformation during coal
combustion, Fuel. 81 (2002) 1251–1258. doi:10.1016/S0016-2361(02)00027-3.
[27] J.C. Hower, K.R. Henke, S. Dai, C.R. Ward, D. French, S. Liu, U.M. Graham, Generation
and nature of coal fly ash and bottom ash, in: T. Robl, A. Oberlink, R. Jones (Eds.), Coal
Combust. Prod., Elsevier, 2017: pp. 21–65. doi:10.1016/B978-0-08-100945-1.00002-2.
[28] S. Srinivasachar, Mineral Behaviour During Coal Combustion 1. Pyrite Transformations,
Prog.Energy Combust.Sci. 16 (1990) 281–292.

of
[29] C.R. Ward, Analysis, Origin and Significance of Mineral Matter in Coal: An Updated
Review, Int. J. Coal Geol. 165 (2016) 1–27. doi:10.1016/j.coal.2016.07.014.
[30] G.L. Fisher, B.A. Prentice, D. Silberman, J.M. Ondov, A.H. Biermann, R.C. Ragaini, A.R.

ro
McFarland, Physical and morphological studies of size-classified coal fly ash, Environ.
Sci. Technol. 12 (1978) 447–451. doi:10.1021/es60140a008.
[31] H. Raclavska, K. Raclavsky, D. Matysek, Colour measurement as a proxy method for

-p
estimation of changes in phase and chemical composition of fly ash formed by
combustion of coal, Fuel. 88 (2009) 2247–2254. doi:10.1016/j.fuel.2009.04.033.
[32] L. Bartonova, Unburned carbon from coal combustion ash: An overview, Fuel Process.
re
Technol. 134 (2015) 136–158. doi:10.1016/j.fuproc.2015.01.028.
[33] S. Wang, Q. Ma, Z.H. Zhu, Characteristics of unburned carbons and their application for
humic acid removal from water, Fuel Process. Technol. 90 (2009) 375–380.
lP

doi:10.1016/j.fuproc.2008.10.010.
[34] M.P. Ketris, Y.E. Yudovich, Estimations of Clarkes for Carbonaceous biolithes: World
averages for trace element contents in black shales and coals, Int. J. Coal Geol. 78 (2009)
135–148. doi:10.1016/j.coal.2009.01.002.
[35] F. Anggara, D.A.Y. Besari, W. Rosita, H.T.B.M. Petrus, The composition and mode of
na

occurrence of rare earth elements and yttrium in fly and bottom ash from coal-fired power
plants in Java, Indonesia, in: S. Dai, Y. Tang (Eds.), Proceeding 35th TSOP, The Society
for Organic Petrology (TSOP), Beijing, 2018: p. 25.
[36] R.L. Thompson, T. Bank, S. Montross, E. Roth, B. Howard, C. Verba, E. Granite,
ur

Analysis of rare earth elements in coal fly ash using laser ablation inductively coupled
plasma mass spectrometry and scanning electron microscopy, Spectrochim. Acta Part B.
143 (2018) 1–11. doi:10.1016/j.sab.2018.02.009.
Jo

[37] A. Kolker, C. Scott, J.C. Hower, J.A. Vazquez, C.L. Lopano, S. Dai, Distribution of rare
earth elements in coal combustion fly ash, determined by SHRIMP-RG ion microprobe,
Int. J. Coal Geol. 184 (2017) 1–10. doi:10.1016/j.coal.2017.10.002.
[38] J.C. Hower, J.G. Groppo, K.R. Henke, U.M. Graham, M.M. Hood, P. Joshi, D. V. Preda,
Ponded and Landfilled Fly Ash as a Source of Rare Earth Elements from a Kentucky
Power Plant, Coal Combustion Gasification Products. 9 (2017) 1–21. doi:10.4177/ccgp-d-
17-00003.1.
[39] J.C. Hower, D. Qian, N.J. Briot, E. Santillan-Jimenez, M.M. Hood, R.K. Taggart, H. Hsu-
Kim, Nano-Scale Rare Earth Distribution in Fly ash Derived from the Combustion of the

19
Fire Clay Coal, Kentucky, Minerals. 9 (2019) 27–32. doi:10.3390/min9040206.
[40] S. Cao, C. Zhou, J. Pan, C. Liu, M. Tang, W. Ji, T. Hu, N. Zhang, Study on Influence
Factors of Leaching of Rare Earth Elements from Coal Fly Ash, Energy Fuels. 32 (2018)
8000–8005. doi:10.1021/acs.energyfuels.8b01316.
[41] P.B. Joshi, D. V. Preda, D.A. Skyler, A. Tsinberg, B.D. Green, W.J. Marinelli, Recovery
of Rare Earth Elements and Compounds From Coal Ash, US Patent No. 8,968,688 B2,
2015. doi:US20130287653 A1.
[42] D. Valeev, I. Kunilova, A. Alpatov, A. Varnavskaya, D. Ju, Magnetite and Carbon
Extraction from Coal Fly Ash Using Magnetic Separation and Flotation Methods,
Minerals. 9 (2019) 5–15. doi:10.3390/min9050320.
[43] J. Liu, S. Dai, X. He, J.C. Hower, T. Sakulpitakphon, Size-Dependent Variations in Fly
Ash Trace Element Chemistry: Examples from a Kentucky Power Plant and with
Emphasis On Rare Earth Elements, Energy Fuels. 31 (2017) 438–447.

of
doi:10.1021/acs.energyfuels.6b02644.
[44] J.C. Hower, J.G. Groppo, U.M. Graham, C.R. Ward, I.J. Kostova, M.M. Maroto-valer, S.
Dai, Coal-Derived Unburned Carbons in Fly Ash : A Review, Int. J. Coal Geol. 179

ro
(2017) 11–27. doi:10.1016/j.coal.2017.05.007.
[45] R.K. Taggart, J.C. Hower, G.S. Dwyer, H. Hsu-Kim, Trends in the Rare Earth Element
Content of U.S.-Based Coal Combustion Fly Ashes, Environ. Sci. Technol. 50 (2016)

-p
5919–5926. doi:10.1021/acs.est.6b00085.
[46] S. Dai, L. Zhao, J.C. Hower, M.N. Johnston, W. Song, P. Wang, S. Zhang, Petrology,
Mineralogy, and Chemistry of Size-Fractioned Fly Ash from The Jungar Power Plant,
re
Inner Mongolia, China, with Emphasis on The Distribution of Rare Earth Elements,
Energy Fuels. 28 (2014) 1502–1514. doi:10.1021/ef402184t.
[47] C. Lanzerstorfer, Pre-processing of coal combustion fly ash by classification for
lP

enrichment of rare earth elements, Energy Reports. 4 (2018) 660–663.


doi:10.1016/j.egyr.2018.10.010.
[48] D. Valeev, A. Mikhailova, A. Atmadzhidi, Kinetics of Iron Extraction from Coal Fly Ash
by Hydrochloric Acid Leaching, Metals (Basel). 8 (2018) 533. doi:10.3390/met8070533.
[49] A. Seidel, Y. Zimmels, Mechanism and kinetics of aluminum and iron leaching from coal
na

fly ash by sulfuric acid, Chem. Eng. Sci. 53 (1998) 3835–3852. doi:10.1016/S0009-
2509(98)00201-2.
[50] U. Kukier, C.F. Ishak, M.E. Sumner, W.P. Miller, Composition and element solubility of
magnetic and non-magnetic fly ash fractions, Environ. Pollut. 123 (2003) 255–266.
ur

doi:10.1016/S0269-7491(02)00376-7.
[51] J.F. King, R.K. Taggart, R.C. Smith, J.C. Hower, H. Hsu-Kim, Aqueous Acid and
Alkaline Extraction of Rare Earth Elements from Coal Combustion Ash, Int. J. Coal Geol.
Jo

195 (2018) 75–83. doi:10.1016/j.coal.2018.05.009.


[52] B.A. Wills, T. Napier-munn, Mineral Processing Technology:An Introduction to the
Practical Aspects of Ore Treatment and Mineral Recovery, 7th ed., Elsevier Science &
Technology Books, 2006. doi:10.1016/B978-075064450-1/50003-5.

20
Jo
ur
na
lP
re
-p
ro
of

21
Table 1. Coal source, combustion temperature, and total REY contents of fly ash samples
obtained from three coal-fired power plants
Tuban Indramayu Paiton-1
Coal source Kalimantan South Sumatera Kalimantan–South Sumatera
mixture
Calorific value (adb, MJ/kg) 22.916 25.845 NA
Combustion temperature (°C) 1400 1388 1100
adb: air-dried basis; NA: not available

of
ro
-p
re
lP
na
ur
Jo

22
Table 2. Elemental compositions of three industrial coal fly ashes.
(Concentrations of major oxide given in %, REY in ppm.)
Element Tuban Indramayu* Paiton-1
SiO2 50.9 42.6 50.6
Al2O3 28.7 20.4 23.6
Fe2O3 9.0 13.8 10.9
CaO 3.3 9.6 5.1
MgO 2.2 6.6 3.8
Na2O 0.8 0.7 0.6
K2O 0.9 0.9 1.6
TiO2 1.0 0.8 1.0
La 42.3 34.5 39.0

of
Ce 84.9 70.0 81.2
Pr 9.7 8.1 9.4
Nd 37.4 32.8 37.0
Sm 7.7 7.1 7.9

ro
Eu 1.7 1.5 1.8
Gd 7.7 6.9 7.3
Tb 1.3 1.2 1.2

-p
Dy 7.8 7.2 7.2
Ho 1.8 1.5 1.5
Er 5.1 4.4 4.5
re
Tm 0.8 0.7 0.6
Yb 5.0 4.5 4.0
Lu 0.8 0.7 0.6
lP

Y 46.4 44.5 41.6


∑ REY 260.4 225.7 244.8
Coutlook 1.1 1.2 1.1
Critical REY (%) 38.3 40.6 38.1
*Major oxide values taken from Nugroho et al. [21]
na
ur
Jo

23
Table 3. Mineral composition (wt.%) determined from XRD analysis
Phase Tuban Indramayu Paiton-1
Amorphous 45 42 47
Spinel 2 5 3
Carbonate* 0 5 0
Quartz 17 20 20
Gypsum 1 0 0
Anhydrite 0 1 1
Mullite 30 12 18
Magnetite 2 9 8

of
Hematite 1 3 0
Rutile 0 0 1
Srebrodolskite 0 3 2
Bayerite 1 0 0

ro
*Carbonates: periclase and calcite

-p
Table 4. Petrographic composition (vol. %) of three industrial coal fly ashes
re
Phase Tuban Indramayu Paiton-1
Unburned carbon 14 13 12
Glass 43 42 40
Quartz 14 15 14
lP

Fe oxides 10 12 9
Spinel 6 6 12
Mullite 14 13 14
na
ur
Jo

24
Table 5. Particle size distribution in each fraction

Particle size (µm) LPSA (%) Dry sieving (%)


> 75 14.4 13
75-63 2.0 2
63-53 2.2 5
53-45 2.4 12
45-38 2.8 14
< 38 76.1 53

of
ro
-p
re
lP
na
ur
Jo

25
Table 6. Chemical compositions of Indramayu coal fly ash as a function of size fraction.
(Major elements are reported in %; REY in ppm.)
Elements > 75 m 75-63 m 63-53 m 53- 45m 45-38 m < 38 m
SiO2 57.7 43.5 45.5 44.3 43.9 39.8
Al2O3 20.0 21.2 20.4 20.7 19.8 21.6
Fe2O3 12.5 12.8 12.5 13.0 13.6 14.5
CaO 2.3 8.3 8.5 9.3 10.4 10.8
MgO 1.4 5.6 5.8 6.3 7.1 7.3
Na2O 0.6 0.7 0.6 0.7 0.6 0.7
K2 O 0.6 0.9 0.8 0.9 0.8 1.0
TiO2 0.7 0.8 0.8 0.8 0.7 0.8
La 24.6 34.3 34.0 35.3 38.2 38.7

of
Ce 48.4 66.7 66.3 69.0 74.1 75.9
Pr 5.3 7.9 8.2 8.5 9.1 9.2
Nd 21.8 31.0 31.5 33.2 34.2 36.4

ro
Sm 3.8 6.3 6.3 7.1 7.9 8.0
Eu 0.9 1.4 1.3 1.5 1.6 1.7
Gd 3.9 7.2 7.1 7.6 7.6 8.2
Tb 0.7 1.2 1.1 1.2 1.2 1.3
Dy
Ho
Er
4.3
0.8
2.7
7.5
1.4
4.5
7.7
1.5
4.4
-p 7.5
1.6
4.7
8.0
1.6
5.0
8.1
1.7
5.1
re
Tm 0.4 0.6 0.7 0.7 0.7 0.8
Yb 2.6 4.2 4.2 4.6 4.9 5.4
Lu 0.4 0.6 0.6 0.7 0.7 0.8
lP

Y 23.2 40.2 40.3 42.9 45.3 47.8


∑REY 143.8 215.0 215.2 226.1 240.1 249.1

Table 7. Enrichment factor (EF) and REY recovery after separation at different magnetic flux
densities
na

Flux Feed to
density magnetic
(G) separation Non-magnetic product Magnetic product
REY Mass REY Mass EF Recovery REY Mass EF Recovery
ur

(ppm) (g) (ppm) (g) (%) (ppm) (g) (%)


10 800 248.98 150 277.14 96.5 1.11 71.63
4300 248.98 150 278.46 122.9 1.12 91.64 175.1 25.5 0.7 11.96
Jo

Table 8. Petrographic composition (vol.%) in Indramayu coal fly ash, size fraction < 38 µm,
nonmagnetic product
Unburned carbon Glass Quartz Fe oxides
16 67 13 3

26

You might also like