You are on page 1of 18

Accepted Manuscript

Title: Green synthesis of mesoporous ␥-Al2 O3 from coal fly


ash with simultaneous on-site utilization of CO2

Authors: Feng Yan, Jianguo Jiang, Nuo Liu, Yuchen Gao,


Yuan Meng, Kaimin Li, Xuejing Chen

PII: S0304-3894(18)30664-2
DOI: https://doi.org/10.1016/j.jhazmat.2018.07.104
Reference: HAZMAT 19613

To appear in: Journal of Hazardous Materials

Received date: 28-7-2017


Revised date: 12-11-2017
Accepted date: 27-7-2018

Please cite this article as: Yan F, Jiang J, Liu N, Gao Y, Meng Y, Li
K, Chen X, Green synthesis of mesoporous ␥-Al2 O3 from coal fly ash with
simultaneous on-site utilization of CO2 , Journal of Hazardous Materials (2018),
https://doi.org/10.1016/j.jhazmat.2018.07.104

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. 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.
Green synthesis of mesoporous γ-Al2O3 from coal fly ash with simultaneous on-site utilization

of CO2

Feng Yan,a Jianguo Jiang,a,b,c * Nuo Liu,a Yuchen Gao,a Yuan Meng,a Kaimin Li,a and Xuejing Chena
a
School of Environment, Tsinghua University, Beijing 100084, China.
b
Key Laboratory for Solid Waste Management and Environment Safety, Ministry of Education, Beijing 100084, China.
c
Collaborative Innovation Center for Regional Environmental Quality, Tsinghua University, Beijing 100084, China.

*Corresponding author:

T
Prof. Dr. Jianguo Jiang (School of Environment, Tsinghua University, Beijing 100084, China.)

IP
Tel./Fax.: +8610-62783548; E-mail address: jianguoj@mail.tsinghua.edu.cn

R
Graphical Abstract

SC
U
N
A
M

Highlights
ED

 Facile conversion of coal fly ash into highly ordered mesoporous γ-Al2O3.
PT

 Practical extraction technique and high extraction efficiency of aluminum.


 Green and rapid synthesis of γ-Al2O3 through the CO2-assistant precipitation.
 Synthetic product shows crystalline framework walls and ordered mesostructure
E
CC

ABSTRACT: Mesoporous Al2O3 with crystalline framework walls has expanded all over the world due to the
various potential applications especially in catalysis. Here, we develop a green and facile approach for the
conversion of coal fly ash (CFA) into ordered mesoporous γ-Al2O3. The practical and promising lime-sinter
A

method was comprehensively studied for the extraction of aluminum from CFA as a first step. The
extraction efficiency of aluminum could reach up to 87.42%, through calcining with CaCO3 at 1390°C for 1 h
and then dissolving in Na2CO3 solution at 70°C for 0.5 h. Combined with the urgent demand for CO2
emission reduction, simulated purified flue gas was introduced to precipitate the Al(OH)3 precursors
without structure-directing agents for just 1 h, followed by calcining at only 400°C or 550°C. A series of
characterizations were conducted to discuss the effect of precipitation temperature and calcination
temperature, resulting the superior product (Al2O3-65/550) with high surface area (230.3 m2 g-1), crystalline

1
γ-Al2O3 phase and ordered mesostructure. This proposed strategy, integrating the on-site recycling of CFA
and utilization of CO2, appears to be promising for scalable production of mesoporous γ-Al2O3.

Keywords: coal fly ash; mesoporous γ-Al2O3; lime-sinter method; CO2-assistant precipitation; aluminum
recovery

1. Introduction
Since the first successful synthesis of mesoporous molecular sieves by Mobil researchers,1,2 the research
of inorganic mesoporous materials has expanded all over the world due to the potential applications in
drug delivery,3 gene delivery,4 biosensors,5 adsorption,6 molecular separation,7 and especially catalysis.8

T
Compared to silica, mesoporous Al2O3 has crystalline framework and intrinsic acid-basic characteristics,9

IP
and thus is more popular as a support for metal catalysts (e.g., Pt,10 Cu,11 Co,12 and Pd13) and bimetal
catalysts (e.g., Cu-Ce,14 Ni-Cu,15 Co-Mo,16 Co-Ni,17 and Mo-Ni18). Since the catalytic performance

R
strongly depends on the properties of catalyst support, there has always been great interest in synthesizing

SC
mesoporous Al2O3 with large surface area, tuned pore size, and well crystalline phase.

Mesoporous Al2O3 is commonly prepared from various aluminum alkoxides,19 aluminate salts,20 and
polyoxocations,21 via a sol-gel self-assembly process followed by calcination.22 To date, significant

U
advances have been achieved to improve the textural properties of mesoporous Al2O3 through the
N
assistance of structure-directing agents (SDAs), including anionic surfactants (e.g., sodium dodecyl
sulfate,23 lauric acid,24 and stearic acid24), cationic surfactants (e.g., C16H33(CH3)3NBr (CTAB)),20 nonionic
A
surfactants (e.g., EO20PO70EO20 (Pluronic P123),19 EO106PO70EO106 (Pluronic F127),19 EO19PO39EO19 (Pluronic
M

P84),21 and EO13PO30EO13 (Pluronic L64)21), and hard templates (e.g., ordered mesoporous carbon).25
However, the mesostructured Al2O3 assembled from molecular precursors are generally comprised of
amorphous framework walls, which lack the stability and surface chemical characteristics of crystalline γ-
ED

Al2O3.9 Although an amorphous alumina phase can be transformed into the γ-Al2O3 phase at a temperature
above 800°C, the calcination under such stringent thermal conditions will inevitably destroy the
mesostructure.26 Zhang et al. synthesized mesostructured γ-Al2O3 via calcining mesostructured
PT

surfactant/boehmite intermediate at only 550°C, and pointed out that crystalline precursor was the key
factor to reduce the phase-transition temperature of γ-Al2O3.27 Similarly, Fulvio et al. synthesized
E

mesoporous γ-Al2O3 directly using commercial boehmite in the presence of P123, which showed better
adsorption properties, higher amount of acidic sites and higher thermal stability in comparison to that
CC

synthesized from aluminum alkoxides.28 Nevertheless, considering the costly aluminum sources and
expensive SDAs, it remains a significant challenge to obtain mesoporous γ-Al2O3 via an environmental-
A

benign process from inexpensive raw materials.

Coal fly ash (CFA) is the major solid waste of coal-fired power plants and its annual generation in China
explosively increased to 620 million tons by 2015;29 yet inappropriate management and irregular
accumulation of CFA makes it a potential pollution source and poses a serious threat to environment.30
Nevertheless, it is noteworthy that the Al2O3 content of CFA generally accounts for 20–50 wt % and even
more than 50 wt % in some special areas of China (e.g., northern Shanxi and western Inner Mongolia),31
making it an extensive, cheap, and readily available source of aluminum. Recently, numerous studies have

2
focused on the extraction of aluminum, which can be classified into sinter methods, acid methods, and
alkaline methods.29 The traditional sinter methods usually couple a high-temperature reaction of CFA with
sintering agent (e.g., lime,32 lime-soda,33,34 and sulfates35–38), followed by an aluminate dissolution
process from sinter-mixture via deionized water or Na2CO3 solution. In order to reduce the energy
consumption, inorganic acid39 and mixed-alkaline solution40 were also proposed for the extraction of
aluminum at lower temperatures; however, the industrial application of these methods is still limited by
the strong corrosion, the highly alkaline residue, and the high-requirement equipment.41 Considering the
abundant feedstock, the mature equipment system,29 and the potential utilization of residues for CO2
capture,42 lime-sinter is still the most promising method for practical application. Although much effort has
been made to increase the extraction efficiency as summarized in Table 1, there are few studies focusing on

T
the further synthesis of mesoporous γ-Al2O3.

IP
In this study, a green and facile approach, including the lime-sinter method and the CO2-assistant

R
precipitation technology, was firstly proposed for the conversion of CFA into mesoporous γ-Al2O3. The
reaction mechanisms of Al2O3 and SiO2 in sintering process were comprehensively discussed, and the

SC
kinetic model was also fitted for the subsequent aluminate dissolution process. Combined with the urgent
demand for carbon emission reduction,43 simulated purified flue gas was introduced into the extracted

U
liquid to obtain the mesoporous γ-Al2O3 without any SDAs. The superior properties of the synthetic
products and the optimal synthesis conditions were investigated via a variety of characterization
N
techniques. Owing to the high extraction efficiency of aluminum and the fast precipitation process with
purified flue gas, this approach could be used for large-scale production of mesoporous γ-Al2O3 from CFA as
A
well as the on-site utilization of CO2.
M

2. Material and Methods


ED

2.1 Extraction of Aluminum from CFA


A typical high-alumina CFA from a coal-fired power plant (Tianjin, China) was used as the aluminum
source to synthesize the mesoporous γ-Al2O3. Based on our previous study,44 the CFA was first desilicated
PT

through alkaline-dissolution process, which was conducted at 110°C for 0.5 h using 25 wt % NaOH solution
(mass ratio of NaOH/FA = 0.5). X-ray fluorescence (XRF, Thermo Scientific, ARL Perform’X, USA, Table S1)
E

analysis indicated that the Al2O3 content of desilicated CFA (DSFA) increased to 46.92 wt % compared with
that of CFA (36.51 wt %), and it would be benefit for the further extraction of aluminum.
CC

5-g DSFA and 10-g CaCO3 powder (99 wt %; Sinopharm, China) were mixed (180 rpm) by a planetary ball
mill (MITR, China) for 10 min, and then the mixture was calcined in a high-temperature muffle furnace
A

(Kejing, KSL-1700X-S, China) with a MoSi2 heating element at different temperatures of 1350–1400°C for
0.5–2 h. The sinter-mixture will automatically disintegrate into fine powder during cooling, due to the
phase transformation (at ~500°C) of Ca2SiO4 accompanied by an 11% volume increase.45 Secondly, 8-g
sinter-mixture was dissolved in 80-mL Na2CO3 solution (0.9 mol/L) at 70°C for 30 min with stirring (300
rpm), and the resultant suspension was immediately filtrated several times with ultrapure water to
separate the residue and the crude aluminate (NaAlO2) solution. The residue was dried at 105°C for further
analysis and was designated based on the calcination condition of sinter-mixture (e.g., “Residue-1390/1h”

3
indicated the calcination at 1390°C for 1 h). The crude NaAlO2 solution was purified by adding Ca(OH)2
suspensions,45 filtrated with ultrapure water, and diluted to 100 mL in a volumetric flask, resulting in the
extracted liquid for synthesis of mesoporous γ-Al2O3. To investigate the kinetic model, the aluminate
dissolution processes were also conducted at different temperatures of 30–70°C for 1–30 min using the
sinter-mixture calcined at 1390ºC for 1 h.

Inductively coupled plasma-atomic emission spectrometry (ICP-AES, Thermo Scientific, iCAP7400, USA)
was used to analyze the concentration of Al (cAl, mg L-1) in the extracted liquid. The extraction efficiency of
aluminum (ηAl) can be calculated with Eq. 1, where MAl2O3 and MAl are the molecular weights of Al2O3 and Al
(g mol-1), respectively; V is the volume of the extracted liquid (mL); mSM is the mass of “sinter-mixture” (g);

T
and WAl2O3 is the weight percentage of Al2O3 in “sinter-mixture”.

IP
6
M  cAl  V  1 0
 Al  A l2 O 3
 100% (Eq. 1)
2 M  m S M  W A l2 O 3

R
Al

SC
2.2 Synthesis of Mesoporous γ-Al2O3
32.9-mL extracted liquid (8,195 mg L-1 Al) derived under the best conditions and 17.1-mL ultrapure water

U
were mixed in a sealed Teflon-lined reactor (BeiLun, China), which had an inlet port at the bottom and an
outlet port on the top. A constant flow of simulated purified flue gas (15 vol % CO2 and 85 vol % N2,43 40 mL
N
min-1) was injected to assist the precipitation of Al(OH)3 (Eq. 2) at 20–80°C for 1 h with stirring (300 rpm).
The Al(OH)3 precursor was then recovered by filtration, washing, and drying at 105°C; the filtrate (Na2CO3
A
solution) could be recycled and reused in the aluminate dissolution process. Finally, the mesoporous γ-
M

Al2O3 was obtained through calcining the Al(OH)3 precursor at 400°C or 550°C for 4 h in air (Eq. 3). The
Al(OH)3 precursors and the synthetic γ-Al2O3 were designated based on their synthesis temperatures (e.g.,
“Al(OH)3-65” for the precipitation temperature of 65°C; “Al2O3-65/400” for the precipitation temperature of
ED

65°C and calcination temperature of 400°C).

2 N a A lO 2 + C O 2  3 H 2 O  N a 2 C O 3 + 2 A l(O H ) 3  (Eq. 2)
PT

400C /550C
2 A l(O H ) 3     
 A l 2 O 3 + 3 H 2 O (Eq. 3)
E
CC

3. Results and Discussion

3.1 Extraction of Aluminum from CFA


A

Fig. 1 presented the influence of calcination time and calcination temperature on the mineral
composition of sinter-mixture based on X-ray diffraction (XRD) patterns. According to our previous study,44
the amorphous silica in “CFA” was removed after desilication, and the scanning electron microscopy (SEM)
images also showed that the smooth surface texture of “CFA” was also replaced by the strip-shaped crystal
grains (Figs. S1a and S1b). The crystalline phases of “DSFA” were mainly mullite (3Al2O3·2SiO2, JCPDS 15-
0776), quartz (SiO2, JCPDS 99-0088), corundum (Al2O3, JCPDS 99-0036), and hydroxysodalite
(Na8Al6Si6O24(OH)2(H2O)2, JCPDS 41-0009). After the calcination of “DSFA” with CaCO3 powder, the above

4
crystalline phases of “DSFA” almost disappeared and new crystalline phases were detected in the sinter-
mixtures, including larnite (β-2CaO·SiO2, JCPDS 33-0302), calcio-olivine (γ-2CaO·SiO2, JCPDS 49-1672),
mayenite (12CaO·7Al2O3, JCPDS 70-2144) and calcium dialuminium oxide (CaO·Al2O3, JCPDS 70-0134).
Larnite was mainly formed via Eqs. S1, S4 and S5 at 1350–1400°C, and then partially transformed into
calcio-olivine at ~500°C during cooling.45 When the calcination time was prolonged and the calcination
temperature was increased, the diffraction peaks of larnite gradually decreased, indicating that this phase
transformation was more thorough at such conditions. Mayenite and calcium dialuminium oxide, formed
from the inactive mullite and corundum (Eqs. S2–S5), could subsequently be dissolved in Na2CO3 solution
for aluminum recycling (Eqs. S6–S7). Mayenite was mainly obtained within shorten time (≤ 1 h, T = 1390°C)
and at lower temperature (≤1390°C, t = 1 h), while calcium dialuminium oxide was mainly obtained for

T
longer time and at higher temperature.

IP
Fig. 2a plotted the extraction efficiencies of aluminum when the sinter-mixtures were calcined under

R
different temperatures and different times. The standard Gibbs free energy changes (ΔGΘ) under different
reaction temperatures were calculated for Eqs. S1–S7 (Table S2), based on the classic thermodynamics

SC
theory and the data in Lang’s Handbook of Chemistry.46,47 Although Eqs. S1–S5 could occur spontaneously
for ΔG < 0 in the range of 1350–1400°C, the reaction rates were considerably influenced by the

U
temperature. Thus, the ηAl rapidly increased as the calcination temperature was increased from 1350°C to
1390°C (t = 0.5 h and 1 h), and achieved the maximum extraction efficiency of 87.42% by the sinter-mixture
N
calcined at 1390ºC for 1 h. However, when the calcination temperature was further increased to 1400°C or
the calcination time was further prolonged to 1.5 h, the aluminate phase in sinter-mixtures was changed
A
from mayenite to calcium dialuminium oxide (Fig. 1). Since calcium dialuminium oxide was not as soluble as
M

mayenite for the smaller absolute value of ΔG,45 the ηAl decreased as the increasing temperature and the
prolonging time. Among the various extraction methods (Table 1), this lime-sinter method has significant
advantages in the cheap feedstock, the low corrosive to equipment and also the competitive extraction
ED

efficiency (0.41 g-Al2O3/g-DSFA), making it a promising method for practical application despite the
relatively high energy consumption.
PT

The chemical properties of residues were also characterized by XRD patterns, XRF analysis and SEM
images. Fig. S2 revealed that the diffraction peaks of larnite and calcio-olivine kept unchanged during the
aluminate dissolution process, indicating that these phases was hardly dissolved in Na2CO3 solution at all.
E

Meanwhile, the diffraction peaks of mayenite and calcium dialuminium oxide almost disappeared, and a
CC

new crystalline phase (calcite, CaCO3, JCPDS 86-0174) was detected in the residues. The change tendency of
Al2O3 content in the residues (from 7.96 to 3.46 wt %, Table S1) was in accordance with the results of ηAl as
the calcination temperature was increased from 1350°C to 1390°C, while the contents of other elements
A

(Si, Fe and Ca) were very similar in all of the residues. Moreover, the morphological contrast of sinter-
mixture and residue was visually showed in SEM images (Figs. S1c and S1d). The sinter-mixture exhibited
various sharp crystal particles in the range of 1–5 μm; the residue mainly contained two kinds of crystal
particles in the range of 1–2 μm, which were identified as CaCO3 (the cubic particle) and Ca2SiO4 (the
clubbed particle) by energy-dispersive spectroscopy (EDS). Significantly, the residue, composed of both
active calcium (calcite) and distributed inert phases (larnite and calcio-olivine), could be further used as an
efficient Ca-based sorbent for CO2 capture.42,43

5
To illustrate the kinetic characteristics of aluminate dissolution process, the change law of ηAl was
studied at different temperatures of 30–70°C for 1–30 min (Fig. 2b). As expected, the ηAl increased from
75.43% to 87.42% (t = 30 min) with the increasing temperature, which would accelerate the reaction rate.
At the dissolution temperature of 70°C, the ηAl rapidly grew to 62.46% in the initial 5 min but slowed down
later, which could thus be divided into two stages of reaction control and diffusion control. Further on, the
classical kinetic model, known as Avrami-Erofeev equation (Eqs. 4 and S8),48 was used to fit this process,
where k is the rate constant; n is a constant; and t is the reaction time, min. Fig. S3a revealed that the
experimental data at all temperatures were well fitted with Avrami-Erofeev equation with high correlation
coefficients (R2 > 0.99), and the kinetic fitting parameters (ln(k), n and R2) were all listed in Table S3. And
then, Arrhenius equation (Eqs. 5 and S9) was applied to evaluate the apparent activation energy (Ea, kJ mol-

T
1
) of this process,48 where A is the frequency factor; R is the ideal gas constant, 8.314 J K-1 mol-1; and T is the

IP
thermodynamic reaction temperature, K. Fig. S3b showed a good linear relationship (R2 = 0.9961) between
ln(k) and T-1 and gave the results of Ea = 19.342 kJ mol-1. Therefore, the apparent reaction kinetic equation

R
for aluminate dissolution process could be expressed as Eq. 6 (n varied from 0.54 to 0.39 as the increasing

SC
temperature).

n
Al
1 exp( k t ) (Eq. 4)

Ea U
N
k A exp( ) (Eq. 5)
RT
A
19342
ln (1 Al
) 485 exp( ) t
n
(Eq. 6)
RT
M
ED
E PT
CC
A

Fig. 1. (a) XRD patterns of “DSFA” and the sinter-mixtures calcined at 1390ºC for 0.5, 1, 1.5, and 2 h; (b) XRD
patterns of the sinter-mixtures calcined for 1 h at 1350ºC, 1360ºC, 1370ºC, 1380ºC, 1390ºC, and 1400ºC.

6
T
Fig. 2. (a) Extraction efficiencies of aluminum when the sinter-mixtures were calcined under different temperatures

IP
of 1350–1400°C and different times of 0.5–2 h (the aluminate dissolution processes were conducted at 70°C for 30
min); (b) extraction efficiencies of aluminum as a function of aluminate dissolution time at different temperatures

R
of 30–70°C.

SC
areagent/dosage,
the reagents and their dosage used for the extraction of aluminum; e.g., “CaCO3 (s)/2.00”
indicated that CaCO3 (s) was used and the mass ratio of CaCO3/DSFA was 2.00. bT, the reaction temperature, ºC.
ct,
U
the reaction time, h. dproduct, the phase of recovery product. eWAl2O3, the weight percentage of Al2O3 in
N
“DSFA” or “CFA”, wt %. fηAl, the extraction efficiency of aluminum, %. gIn this study. hThe mass ratio of
reagent/DSFA or reagent/CFA. iThe liquid to solid ratio (L/S) of reagent/DSFA or reagent/CFA. jThe weight
A
percentage of Al2O3 in “DSFA”. kN. A., not available. lN. N., not need (this method only need step 1).
M

3.2 Characterization of Mesoporous γ-Al2O3


ED

The XRD patterns of Al(OH)3 precursors obtained at different precipitation temperatures were shown in
Fig. 3a. Precipitation at 20°C gave rise to an amorphous structure, in spite of the very weak diffraction
peaks referred to aluminum oxide hydroxide (AlO(OH), JCPDS 49-0133) and bayerite (Al(OH)3, JCPDS 74-
PT

1119). And then the crystallinity of Al(OH)3 precursors gradually increased with the increasing precipitation
temperature, which was indicated by the stronger diffraction peaks of bayerite (2θ = 18.813, 20.297 and
40.652°). When the precipitation temperature was raised to 80°C, another crystalline phase of gibbsite
E

(Al(OH)3, JCPDS 76-1782) was also formed in “Al(OH)3-80”. Fig. S4 presented the N2 temperature-
CC

programmed decomposition (N2-TPD) analysis of Al(OH)3 precursors. All the Al(OH)3 precursors showed a
total weight loss of ~34.6%, which was consistent with theoretical analysis (Eq. 3). The well crystalline
precursors showed a concentrated decomposition temperature region of 210–260°C, while “Al(OH)3-20”
A

and “Al(OH)3-35” gradually decomposed over the temperature range of 25–400°C. Significantly, the
endothermic peak in heatflow curve, associated with the decomposition of Al(OH)3, enlarged and shifted
from 235 to 253°C as the increasing crystallinity.

The XRD patterns of synthetic γ-Al2O3 calcined at 400°C and 550°C were shown in Figs. 3b and 3d,
respectively. Generally, the crystalline phase of γ-Al2O3 (JCPDS 10-0425) was formed at temperature ≥
800°C,19 thus an amorphous structure was found in both “Al2O3-20/400” and “Al2O3-20/550”. However, the
amorphous Al2O3 phase was gradually converted to γ-Al2O3 phase when the precipitation temperature for

7
Al(OH)3 precursors was increased, which was obviously confirmed by the stronger diffraction peaks at 2θ =
37.603, 45.862 and 67.032°. This inspiring result could be attributed to the formation of crystalline Al(OH)3
precursors (bayerite or gibbsite), which was responsible to reduce the phase-transition temperature of γ-
Al2O3.27 Besides, the crystallinity of γ-Al2O3 also slightly increased as the calcination temperature increased
from 400 to 550°C. Commonly, at least seven transition aluminas could possibly form during the calcination
of Al(OH)3 or AlO(OH), among which γ-Al2O3 is perhaps the most extensively used in catalytic process.21
Therefore, the synthesis of γ-Al2O3 at such low temperature (400°C) is of great importance, which could not
only provide the crystalline framework but also preserve the mesoporous structure.

The small-angle XRD (SAXRD) patterns of synthetic γ-Al2O3 calcined at 400°C were shown in Fig. S5b. The

T
absence of any diffraction peak in “Al2O3-20/400”, “Al2O3-35/400” and “Al2O3-50/400” indicated the

IP
absence of any mesostructured,19 which was also confirmed with pore size distributions. When the
precipitation temperature for Al(OH)3 precursors increased, a weak diffraction peak associated with the (1

R
0 0) reflection was observed in both “Al2O3-65/400” and “Al2O3-80/400”, which could be ascribed to the
formation of p6mm hexagonal symmetry.18 For another, sharp diffraction peaks of (1 0 0) reflection were

SC
observed in all synthetic γ-Al2O3 calcined at 550°C (Fig. 3c), in agreement with pore size distributions. The
location of the diffraction peaks (2θ100) and the corresponding interplanar spacing (d100) were calculated

U
and summarized in Table 2. It was noticeable that the diffraction peaks shifted to higher degrees as the
increasing precipitation temperature, which implied that “Al2O3-65/550” and “Al2O3-80/550” exhibited
N
larger d-spacing values.21
A
M
ED
E PT
CC
A

Fig. 3. (a) XRD patterns of “Al(OH)3-20”, “Al(OH)3-35”, “Al(OH)3-50”, “Al(OH)3-65”, and “Al(OH)3-80”; (b)

8
XRD patterns of “Al2O3-20/400”, “Al2O3-35/400”, “Al2O3-50/400”, “Al2O3-65/400”, and “Al2O3-80/400”; (c)
SAXRD patterns and (d) XRD patterns of “Al2O3-20/550”, “Al2O3-35/550”, “Al2O3-50/550”, “Al2O3-65/550”, and
“Al2O3-80/550”.

The N2 physisorption isotherms were shown in Figs. 4a and 4c for synthetic γ-Al2O3 calcined at 400°C and
550°C, respectively. “Al2O3-20/400” and “Al2O3-35/400” possessed type II physisorption isotherms with a
H3 hysteresis loop at P·P0-1 = 0.7–1.0, caused by capillary condensation of nitrogen in slit-shaped pores
between the plate-like particles.44 With the increasing precipitation temperature, the physisorption
isotherms of “Al2O3-65/400” and “Al2O3-80/400” changed to a type IV mode and exhibited a distinct
condensation step at P·P0-1 = 0.45–0.55, which was typical for uniform mesopores.18 And the H1 hysteresis

T
loop at P·P0-1 = 0.45–1.0 suggested the approximately uniform spheres in fairly regular array and hence to

IP
have narrow distributions of pore size.18 The similar change tendency, as the increasing precipitation
temperature, was also observed for synthetic γ-Al2O3 calcined at 550°C. Notably, “Al2O3-65/550” and

R
“Al2O3-80/550” possessed steeper condensation steps compared to those calcined at 400°C, which

SC
indicated a higher degree of uniform mesostructure.

Fig. 4b exhibited the corresponding pore size distributions of synthetic γ-Al2O3 calcined at 400°C.

U
Although “Al2O3-20/400” had a very wide pore size distribution of 1.7–100 nm, higher precipitation
temperature gave rise to a narrow pore size distribution centered at 1.7–5.5 nm for “Al2O3-35/400” and
N
“Al2O3-50/400” and even a narrower pore size distribution centered at 1.7–3.8 nm for “Al2O3-65/400” and
“Al2O3-80/400”. When the calcination temperature was increased to 550°C, all synthetic γ-Al2O3 showed a
A
relatively narrow pore size distribution (Fig. 4d), in accordance with the SAXRD results. Two relatively
M

concentrated peaks centered at ~2.2 nm and ~15 nm were observed for both “Al2O3-20/550” and “Al2O3-
35/550”, while only one sharper concentrated peak was observed for the others, indicating the more
uniform mesostructure of “Al2O3-65/550” and “Al2O3-80/550”.
ED

The structural parameters, such as specific surface area (SBET), total pore volume (Vpore), and average
pore size (DBJH), were summarized in Table 2. With the increasing precipitation temperature, the SBET and
PT

Vpore of Al(OH)3 precursors rapidly decreased from 230.5 m2 g-1 and 0.74 cm3 g-1 to 4.0 m2 g-1 and 0.009 cm3
g-1, respectively, which could be ascribed to the transformation from amorphous phase to crystalline phase.
However, The SBET of synthetic γ-Al2O3 calcined at 400°C significantly increased with the increasing
E

precipitation temperature, reaching a maximum of 404.3 m2 g-1 for “Al2O3-65/400” and slightly decreasing
CC

to 383.1 m2 g-1 for “Al2O3-80/400”. On the contrary, the Vpore and DBJH of synthetic γ-Al2O3 decreased with
the increasing precipitation temperature. As expected, the higher calcination temperature would
decreased the SBET of synthetic γ-Al2O3, yet, “Al2O3-65/550” still exhibited a high SBET of 230.3 m2 g-1.
A

Inspiringly, the large SBET and narrow pore size distribution combined with crystalline walls, which could
even be comparable to the material synthesized with SDAs,19 remarkably promoted the potential
applications of these mesoporous γ-Al2O3 in catalysis.

9
T
R IP
SC
U
N
A
Fig. 4. (a) N2 physisorption isotherms and (b) pore size distributions of “Al2O3-20/400”, “Al2O3-35/400”, “Al2O3-
M

50/400”, “Al2O3-65/400”, and “Al2O3-80/400”; (c) N2 physisorption isotherms and (d) pore size distributions of
“Al2O3-20/550”, “Al2O3-35/550”, “Al2O3-50/550”, “Al2O3-65/550”, and “Al2O3-80/550”.
ED

aS
BET, specific surface area. bVpore, total pore volume. cDBJH, average pore size. d2θ100, the location of diffraction
peaks for the (1 0 0) reflection. ed100, interplanar spacing of the (1 0 0) reflection, which was calculated by the
Bragg’s Law, 2  d 100  s in  100   ; fN.A., not available;
PT

Fig. S7 showed the changes in morphology of the Al(OH)3 precursors and the synthetic γ-Al2O3 with
E

increasing precipitation temperature. “Al(OH)3-20” showed uniform ellipsoid-like agglomerates in the range
of 0.5–1 μm, which consisted of primary particles of ~100 nm. As the increasing precipitation temperature,
CC

the size of agglomerates gradually enlarged to 1–2 μm, and the intergranular gaps between primary
particles also basically disappeared. Instead, angular agglomerates of crystalline phase were observed in
“Al(OH)3-50”, “Al(OH)3-65” and “Al(OH)3-80”, in accordance with the XRD results. After calcining the Al(OH)3
A

precursors at 400°C or 550°C, the more spherical agglomerates and more distinct intergranular gaps were
observed in the synthetic γ-Al2O3, due to the decomposition of Al(OH)3. Although the amorphous phase
gradually transformed to γ-Al2O3 phase with the increasing precipitation temperature, the crystalline
agglomerates still maintained intergranular gaps between primary particles, which was also confirmed by
pore size distributions.

10
The textural structures of “Al2O3-65/400” and “Al2O3-65/550” were obtained from transmission electron
microscopy (TEM) analysis. “Al2O3-65/400” showed plate-like agglomerates with a mostly disordered
mesoporous structure, except for a small fraction of ordered mesopores (Fig. 5b); while “Al2O3-65/550”
clearly showed a long-range ordering of the mesopores with a width of ~4.6 nm (Fig. 5e), close to the
parameters of DBJH and d100. The selected-area electron diffraction (SAED) pattern (insets of Figs. 5b and 5e)
of the ordered mesostructure domains confirmed the mesoporous wall was crystalline with γ-Al2O3 phase
(cubic, space group Fd-3m(227)), in line with the XRD patterns of “Al2O3-65/400” and “Al2O3-65/550”.
Additional proof for the crystallinity of the framework walls was given by high-resolution TEM (HRTEM,
Figs. 5c and 5f) investigations, which revealed well-defined lattices recorded from [001] orientation and
embedded mesoporous networks. The insets of HRTEM images showed fast Fourier transform (FFT)

T
diffractograms also confirming the crystalline γ-Al2O3 phase.

R IP
SC
U
N
A
M
ED

Fig. 5. SEM images of (a) “Al2O3-65/400” and (d) “Al2O3-65/550”; TEM images and the SAED patterns (inset) of
PT

(b) “Al2O3-65/400” and (e) “Al2O3-65/550”; HRTEM images and the FFT diffractograms (inset) of (c) “Al2O3-
65/400” and (f) “Al2O3-65/550”.
E

4. Conclusion
CC

An environmentally friendly strategy was developed to convert coal fly ash, a cheap, abundant, and
aluminum-rich industrial waste, into mesoporous γ-Al2O3. The lime-sinter method was comprehensively
A

studied for the extraction of aluminum from CFA as a first step, achieving a high extraction efficiency of
87.42% through calcining with CaCO3 at 1390°C for 1 h and then dissolving in Na2CO3 solution at 70°C for
0.5 h. Generally, synthesis of mesoporous γ-Al2O3 requires a long aging time (> 24 h) for particle growth and
the guidance of expensive SDAs.22 However, the products in this study could be prepared without SDAs in
just 1 h and also showed an ordered mesostructure with crystalline framework walls, owing to the
precipitation of Al(OH)3 precursors at 65–80°C with the assistance of alternative precipitant (purified flue

11
gas). This inspiring result thus provides a practical and promising approach to mesoporous γ-Al2O3 from
CFA, as well as utilizing the emitted CO2 from industrial point source.

Acknowledgments
We gratefully acknowledge the National Natural Science Foundation of China (grant no. 21576156) and
the Tsinghua University Initiative Scientific Research Program (grant no. 2014z22075). We also thank the
support of the Beijing National Center for Electron Microscopy at Tsinghua University and the support from
Shanghai Tongji Gao Tingyao Environmental Science and Technology Development Foundation.

T
Reference

IP
References

(1) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Ordered mesoporous molecular-sieves synthesized by a liquid-crystal template mechanism.

R
Nature 1992 359
, (6397), 710–712.

SC
(2) Beck, J. S.; Vartuli, J. C.; Roth, W. J.; Leonowicz, M. E.; Kresge, C. T.; Schmitt, K. D.; Chu, C. T. W.; Olson, D. H.; Sheppard, E. W.; McCullen, S. B.; Higgins, J. B.;
Schlenker, J. L. A new family of mesoporous molecular- sieves prepared with liquid-crystal templates. J. Am. Chem. Soc. 1992, 114 (
27), 10834–10843.

U
(3) Ashley, C. E.; Carnes, E. C.; Phillips, G. K. The targeted delivery of multicomponent cargos to cancer cells
N
by nanoporous particle-supported lipid bilayers. Nat. Mater. 2011, 10 (5), 389–397.
A
(4) Kim, M. H.; Na, H. K.; Kim, Y. K.; Ryoo, S. R.; Cho, H. S.; Lee, K. E.; Jeon, H.; Ryoo, R. ; Min, D. H. Facile synthesis of monodispersed mesoporous silica nanoparticles
with ultralarge pores and their application in gene delivery. ACS Nano 2011, 5
(5), 3568–3576.
M

(5) Carrasquilla, C.; Lau, P. S.; Li, Y. F.; Brennan, J. D. Stabilizing structure-switching signaling RNA aptamers by entrapment in sol-gel derived materials for solid-phase
assays. J. Am. Chem. Soc. 2012, 134
(26), 10998–11005.
ED

(6) Patra, A. K.; Dutta, A.; Bhaumik, A. Self-assembled mesoporous γ-Al2O3 spherical nanoparticles and their
efficiency for the removal of arsenic from water. J. Hazard. Mater. 2012, 201, 170–177.
PT

(7) Zhao, W. R.; Tang, Y.; Wan, Y. P.; Li, L.; Yao, S.; Li, X. W.; Gu, J. L.; Li, Y. S.; Shi, J. L. Promotion effects of SiO2
or/and Al2O3 doped CeO2/TiO2 catalysts for selective catalytic reduction of NO by NH3. J. Hazard. Mater. 2014,
278, 350–359.
E
CC

(8) Wang, L.; He, H.; Zhang, C. B.; Wang, Y. F.; Zhang, B. Effects of precursors for manganese-loaded gamma-
Al2O3 catalysts on plasma-catalytic removal of o-xylene. Chem. Eng. J. 2016, 288, 406–413.

(9) Rascón, F.; Wischert, R.; Copéeret C. Molecular nature of support effects in single-site heterogeneous
A

catalysts: silica vs. alumina. Chem. Sci. 2011, 2 (8), 1449–1456.

(10) Adibi, P. T. Z.; Pingel, T.; Olsson, E.; Grönbeck, H.; Langhammer, C. Plasmonic nanospectroscopy of platinum
catalyst nanoparticle sintering in a mesoporous alumina support. ACS Nano 2016, 10 (5), 5063–5069.

(11) Ham, H.; Kim, J.; Cho, S. J.; Choi, J. H.; Moon, D. J.; Bae, J. W. Enhanced stability of spatially confined copper
nanoparticles in an ordered mesoporous alumina for dimethyl ether synthesis from syngas. ACS Catal. 2016, 6
(9), 5629–5640.

12
(12) Tsakoumis, N. E.; Walmsley, J. C.; Rønning, M.; Beek, W. V.; Rytter, E.; Holmen, A. Evaluation of reoxidation
thresholds for γ-Al2O3-supported cobalt catalysts under fischer-tropsch synthesis conditions. J. Am. Chem. Soc.
2017, 139 (10), 3706–3715.

(13) Wang, B. W.; Fang, W. W.; Si, L. L.; Li, Y.; Yan, X. L.; Chen, L. G.; Wang, S. T. Construction of 2-(2’-Hydroxy-5’-
methylphenyl)benzotriazole over Pd/γ-Al2O3 by a continuous process. ACS Sustainable Chem. Eng. 2015, 3 (8),
1890–1896.

(14) Chen, F. T.; Yu, S. C.; Dong, X. P.; Zhang, L.; Wu, Q. F.; Preparation and characterization of PbO2 electrode
and its application in electro-catalytic degradation of o-aminophenol in aqueous solution assisted by CuO–
Ce2O3/γ-Al2O3 catalyst. J. Hazard. Mater. 2013, 260, 747–753.

T
(15) Mansouri, A.; Khodadadi, A. A. Mortazavi, Y. Ultra-deep adsorptive desulfurization of a model diesel fuel on

IP
regenerable Ni–Cu/γ-Al2O3 at low temperatures in absence of hydrogen. J. Hazard. Mater. 2014, 271, 120–130.

R
(16) Yang, Y.; Xu, S. M.; Li, Z.; Wang, J. L.; Zhao, Z. W.; Xu Z. H. Oil removal of spent hydrotreating catalyst

SC
CoMo/Al2O3 via a facile method with enhanced metal recovery. J. Hazard. Mater. 2016, 318, 723–731.

(17) Arbag, H.; Yasyerli, S.; Yasyerli, N.; Dogu, G.; Dogu, T. Enhancement of catalytic performance of Ni based
mesoporous alumina by Co incorporation in conversion of biogas to synthesis gas. Appl. Catal. B: Environ. 2016,
198, 254–265.
U
N
(18) Liu, H.; Li, Y. P.; Yin, C. L.; Wu, Y. L.; Chai, Y. M.; Dong, D. M.; Li, X. H.; Liu, C. G. One-pot synthesis of ordered
A
mesoporous NiMo-Al2O3 catalysts for dibenzothiophene hydrodesulfurization. Appl. Catal. B: Environ. 2016, 198,
493–507.
M

(19) Yuan, Q.; Yin, A. X.; Luo, C.; Sun, L. D.; Zhang, Y. W.; Duan, W. T.; Liu, H. C.; Yan, C. H. Facile synthesis for
ordered mesoporous γ-aluminas with high thermal stability. J. Am. Chem. Soc. 2008, 130 (11), 3465–3472.
ED

(20) Wu, W.; Wan, Z. J.; Zhu, M. M.; Zhang, D. K. A facile route to aqueous phase synthesis of mesoporous
alumina with controllable structural properties. Micropor. Mesopor. Mater. 2016, 223, 203–212.
PT

(21) Zhang, Z. R.; Pinnavaia T. J. Mesostructured γ-Al2O3 with a lathlike framework morphology. J. Am. Chem.
Soc. 2002, 124 (41), 12294–12301.
E

(22) Márquez-Alvarez, C.; Žilková, N.; Pérez-Pariente, J.; Čejka, J. Synthesis, characterization and catalytic
CC

applications of organized mesoporous aluminas. Catal. Rev. 2008, 50 (2), 222–286.

(23) Wang, W. W.; Zhou, J. B. Zhang, Z.; Yu, J. G.; Cai, W. Q. Different surfactants-assisted hydrothermal synthesis
of hierarchical γ-Al2O3 and its adsorption performances for parachlorophenol. Chem. Eng. J. 2013, 233, 168–175.
A

(24) Vaudry, F.; Khodabandeh, S.; Davis, M. E. Synthesis of pure alumina mesoporous materials. Chem. Mater.
1996, 8 (7), 1451–1464.

(25) Wu, Z. X.; Li, Q.; Feng, D.; Webley, P. A.; Zhao, D. Y. Ordered mesoporous crystalline γ-Al2O3 with variable
architecture and porosity from a single hard template. J. Am. Chem. Soc. 2010, 132 (34), 12042–12050.

(26) Zhang, Z. R.; Pinnavaia, T. J. Mesostructured forms of the transition phases η- and χ-Al2O3. Angew. Chem.
Int. Ed. 2008, 47 (39), 7501–7504.

13
(27) Zhang, Z. R.; Hicks, R. W.; Pauly, T. R. Pinnavaia, T. J. Mesostructured Forms of γ-Al2O3. J. Am. Chem. Soc.
2002, 124 (8), 1592–1593.

(28) Fulvio, P. F.; Brosey, R. I.; Jaroniec, M. Synthesis of mesoporous alumina from boehmite in the presence of
triblock copolymer. ACS Appl. Mater. Inter. 2010, 2 (2), 588–593.

(29) Sun, L. Y.; Luo, K; Fan, J. R.; Lu, H. L. Experimental study of extracting alumina from coal fly ash using
fluidized beds at high temperature. Fuel 2017, 199, 22–27.

(30) Zhang, L. G.; Xu, Z. M. An environmentally-friendly vacuum reduction metallurgical process to recover
germanium from coal fly ash. J. Hazard. Mater. 2016, 312, 28–36.

T
(31) Qi, L. Q.; Yuan, Y. T. Characteristics and the behavior in electrostatic precipitators of high-alumina coal fly

IP
ash from the Jungar power plant, Inner Mongolia, China. J. Hazard. Mater. 2011, 192, 222–225.

(32) Grzymek, J.; Derdackagrzymek, A.; Konik, Z.; Stok, A.; Iwanciw, J. The new way of alumina lixiviation from

R
sinters containing 12CaO·7Al2O3 in J. Grzymek’s method [M]. Light Metals 1988, 129–133.

SC
(33) Bai, G. H.; Teng, W.; Wang, X. G.; Qin, J. G.; Xu, P.; Li, P. C. Alkali desilicated coal fly ash as substitute of
bauxite in lime-soda sintering process for aluminum production. Trans. Nonferrous Met. Soc. China 2010, 20 (S1),
s169–s175.
U
N
(34) Jiang, Z. Q.; Ma, H. W.; Yang, J.; MA, X.; Yuan, J. Y. Thermal decomposition mechanism of desilication coal fly
ash by low-lime sinter method for alumina extraction. Ferroelectrics 2015, 486 (1), 143–155.
A
(35) Guo, C. B.; Zou, J. J.; Wei, C. D.; Jiang, Y. S. Comparative study on extracting alumina from circulating
M

fluidized-bed and pulverized-coal fly ashes through salt activation. Energy Fuels 2013, 27 (12), 7868–7875.

(36) Guo, C. B.; Zou, J. J.; Jiang, Y. S.; Huang, T. P.; Cheng, Y.; Wei, C. D. Thermal activation of coal fly ash by
ED

sodium hydrogen sulfate for alumina extraction. J. Mater. Sci. 2014, 49 (12), 4315–4322.

(37) Wang, R. C.; Zhai, Y. C.; Wu, X. W.; Ning, Z. Q.; Ma, P. H. Extraction of alumina from fly ash by ammonium
PT

hydrogen sulfate roasting technology. Trans. Nonferrous Met. Soc. China 2014, 24 (5), 1596–1603.

(38) Wu, Y. S.; Xu, P.; Chen, J.; Li, L. S; Li, M. C. Effect of temperature on phase and alumina extraction efficiency
of the product from sintering coal fly ash with ammonium sulfate. Chinese J. Chem. Eng. 2014, 22 (11-12), 1363–
E

1367.
CC

(39) Nayak, N.; Panda, C. R. Aluminium extraction and leaching characteristics of Talcher Thermal Power Station
fly ash with sulphuric acid. Fuel 2010, 89 (1), 53–58.
A

(40) Su, S. Q.; Yang, J.; Ma, H. W.; Jiang, F.; Liu, Y. Q.; Li, G. Preparation of ultrafine aluminum hydroxide from
coal fly ash by alkali dissolution process. Integr. Ferroelectr. 2011, 128, 155–162.

(41) Wang, Z. H.; Ma, S. H.; Tang, Z. H.; Wang, X. H.; Zheng, S. L. Effects of particle size and coating on
decomposition of alumina-extracted residue from high-alumina fly ash. J. Hazard. Mater. 2016, 308, 253–263.

14
(42) Yan, F.; Jiang, J. G.; Li, K. M.; Tian, S. C.; Zhao, M.; Chen, X. J. Performance of coal fly ash stabilized,
CaO-based sorbents under different carbonation-calcination conditions. ACS Sustainable Chem. Eng. 2015,
3 (9), 2092–2099.

(43) Yan, F.; Jiang, J. G.; Li, K. M.; Liu, N.; Chen, X. J.; Gao, Y. C.; Tian, S. C. Green synthesis of nanosilica from
coal fly ash and its stabilizing effect on CaO sorbents for CO2 capture. Environ. Sci. Technol. 2017, 51 (13),
7606–7615.

(44) Yan, F.; Jiang, J. G.; Tian, S. C.; Liu, Z. W.; Shi, J.; Li, K. M.; Chen, X. J.; Xu, Y. W. A green and facile synthesis of
ordered mesoporous nanosilica using coal fly ash. ACS Sustainable Chem. Eng. 2016, 4 (9), 4654–4661.

T
(45) Yao, Z. T.; Xia, M. S.; Sarker, P. K.; Chen, T. A review of the alumina recovery from coal fly ash, with a focus in
China. Fuel 2014, 120, 74–85.

IP
(46) Speight, J. G. Lang's handbook of chemistry. McGraw-Hill Professional Publishing, 15th edition, 1998,

R
Chapter 6, pp 81–123.

SC
(47) Barin, I.; Platzki, G. Therrnochernical data of pure substances. Wiley-VCH Publishing, New York, NY, USA,
third edition, 1995, Chapter 12, pp 1–1885.

U
(48) Padilla, R.; Sohn, H. Y. Sintering kinetics and alumina yield in lime-soda sinter process for alumina from coal
wastes. Metall. Mater. Trans. B 1985, 16 (2), 385–395.
N
(49) Bai, G. H.; Qiao, Y. H.; Shen, B.; Chen, S. L. Thermal decomposition of coal fly ash by concentrated sulfuric
A
acid and alumina extraction process based on it. Fuel Process. Technol. 2011, 92 (6), 1213–1219.
M

(50) Wu, C. Y.; Yu, H. F.; Zhang, H. F. Extraction of aluminum by pressure acid-leaching method from coal fly ash.
Trans. Nonferrous Met. Soc. China 2012, 22 (9), 2282–2288.
ED

(51) Tripathy, A. K.; Sarangi, C. K.; Tripathy, B. C.; Sanjay, K.; Bhattacharya, I. N.; Mahapatra, B.K.; Behera, P. K.;
Satpathy, B. K. Aluminium recovery from NALCO fly ash by acid digestion in the presence of fluoride ion. Int. J.
Miner. Process. 2015, 138, 44–48.
PT

(52) Xu, D. H.; Li, H. Q.; Bao, W. J.; Wang, C. Y. A new process of extracting alumina from high-alumina coal fly
ash in NH4HSO4+H2SO4 mixed solution. Hydrometallurgy 2016, 165 (SI) 336–344.
E

(53) Guo, Y. X.; Li, Y. Y.; Cheng, F. Q.; Wang, M.; Wang, X. M. Role of additives in improved thermal activation of
CC

coal fly ash for alumina extraction. Fuel Process. Technol. 2013, 110, 114–121.

(54) Guo, Y. X.; Zhao, Z. S.; Zhao, Q.; Cheng, F. Q. Novel process of alumina extraction from coal fly ash by pre-
A

desilicating—Na2CO3 activation—Acid leaching technique. Hydrometallurgy 2017, 169, 418–425.

(55) Yang, Q. C.; Ma, S. H.; Zheng, S. L.; Zhang, R. Recovery of alumina from circulating fluidized bed combustion
Al-rich fly ash using mild hydrochemical process. Trans. Nonferrous Met. Soc. China 2014, 24 (4), 1187–1195.

(56) Li, H. Q.; Hui, J. B.; Wang, C. Y.; Bao, W. J.; Sun, Z. H. Extraction of alumina from coal fly ash by mixed-
alkaline hydrothermal method. Hydrometallurgy 2014, 147, 183–187.

15
(57) Ding, J.; Ma, S. H.; Zheng, S. L.; Zhang, Y.; Xie, Z. L.; Shen, S.; Liu, Z. K. Study of extracting alumina from high-
alumina PC fly ash by a hydro-chemical process. Hydrometallurgy 2016, 161, 58–64.

(58) Sun, Z. H.; Li, H. Q.; Bao, W. J.; Wang, C. Y. Mineral phase transition of desilicated high alumina fly ash with
alumina extraction in mixed alkali solution. Int. J. Miner. Process. 2016, 153, 109–117.

T
Table 1

IP
Comparison of the extraction methods and efficiencies of aluminum from CFA in different studies.

step 1 step 2

R
productd WAl2O3e ηAlf
method reagent Tb tc reagent T t (wt %) (%)

SC
/dosagea (ºC) (h) /dosage (ºC) (h)
0.9 mol/L
sinterg CaCO3(s)/2.00h 1390 1 70 0.5 NaAlO2 46.92j 87.42
Na2CO3(aq)/10i

sinter33
Na2CO3(s)/0.51h
+ CaCO3(s)/1.01h
1200 1
1.5 mol/L
NaOH(aq)/4i U 85 0.17 NaAlO2 49.20j 92.30
N
Na2CO3(s)/1.02h 0.5 mol/L
sinter34 1050 2 90 0.5 NaAlO2 55.15j 94.50
+ CaCO3(s)/0.42h NaOH(aq)/15i
A
sinter35 Na2S2O7(s)/2.40h 420 2 H2O/N.A.k N.A. N.A. Na3Al(SO4)3 51.70 93.30
M

sinter36 NaHSO4(s)/3.65h 500 2 H2O/3i 100 0.5 Na3Al(SO4)3 51.70 90.00


sinter37 NH4HSO4(s)/3.48h 400 0.75 H2O/N.A. N.A. N.A. NH4Al(SO4)2 38.53 90.11
ED

sinter38 (NH4)2SO4(s)/2.85h 400 3 H2O/5i 90 3 NH4Al(SO4)2 40.70 85.38


acid39 18 mol/L H2SO4(aq)/2i 200 4 H2O/5i 90 N.A. Al2(SO4)3 29.09 84.17
acid49 98 wt % H2SO4(l)/2h 275 2 H2O/6i 85 0.5 Al2(SO4)3 43.87 87.00
PT

acid50 50 wt % H2SO4(aq)/4i 180 4 H2O/N.A. 90 N.A. Al2(SO4)3 37.70 82.40


25 wt % H2SO4 +
acid51 90 3 N. Nl N. N N. N Al2(SO4)3 24.00 92.00
E

1.5 wt % HF(aq)/10i
25 wt % HCl +
CC

acid51 90 3 N. N N. N N. N AlCl3 24.00 60.00


1.5 wt % HF(aq)/10i
5 mol/L NH4HSO4 +
acid52 220 4 H2O/N.A. 90 N.A. NH4Al(SO4)2 50.45 87.80
5 mol/L H2SO4(aq)/2.5i
A

Na2CO3(s)/0.5h 20 wt %
acid53 700 2 100 2 AlCl3 33.59 95.00
+ NaOH(s)/0.5h HCl(aq)/3i
20 wt %
acid54 Na2CO3(s)/0.55h 850 2 100 2 AlCl3 35.40 87.00
HCl(aq)/5i
Ca(OH)2(s)/0.43h +
alkaline 40 260 1 N. N N. N N. N NaAlO2 33.61j 89.00
20 mol/L NaOH(aq)/3i
Ca(OH)2(s)/0.52h +
alkaline 55 280 1 N. N N. N N. N NaAlO2 32.43 92.31
45 wt % NaOH(aq)/9i

16
Ca(OH)2(s)/0.50h +
alkaline 56 260 0.75 N. N N. N N. N NaAlO2 48.80 91.30
40 wt % NaOH(aq)/12i
Ca(OH)2(s)/0.55h +
alkaline 57 280 2 N. N N. N N. N NaAlO2 49.50 96.03
45 wt % NaOH(aq)/13i
Ca(OH)2(s)/0.36h +
alkaline 58 280 2 N. N N. N N. N NaAlO2 50.73j 92.00
50 wt % NaOH(aq)/6i

Table 2

T
Structural characteristics of the Al(OH)3 precursors and the synthetic mesoporous γ-Al2O3.

IP
SBETa Vporeb DBJHc 2θ100d d100e
sample
(m2 g-1) (cm3 g-1) (nm) (º) (Å)

R
Al(OH)3-20 230.5 0.74 15.04 N.A.f N.A.

SC
Al(OH)3-35 142.6 0.31 9.72 N.A. N.A.
Al(OH)3-50 9.2 0.021 36.21 N.A. N.A.
Al(OH)3-65 7.4 0.011
U
36.39 N.A. N.A.
N
Al(OH)3-80 4.0 0.009 43.03 N.A. N.A.

Al2O3-20/400 196.1 0.82 15.22 N.A. N.A.


A
Al2O3-35/400 257.0 0.43 8.28 N.A. N.A.
M

Al2O3-50/400 316.4 0.28 3.97 N.A. N.A.


Al2O3-65/400 404.3 0.23 2.83 1.58 55.9
ED

Al2O3-80/400 383.1 0.25 2.77 1.56 56.6

Al2O3-20/550 166.7 0.51 10.94 0.88 100.4


PT

Al2O3-35/550 207.7 0.42 7.46 0.98 90.1


Al2O3-50/550 236.1 0.41 5.78 1.02 86.5
Al2O3-65/550 230.3 0.25 3.84 1.24 71.2
E

Al2O3-80/550 227.4 0.25 3.91 1.18 74.8


CC
A

17

You might also like