Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Accepted Manuscript

Hydrothermal synthesis of high purity zeolite A from natural kaolin without


calcination

Jing-Quan Wang, Ya-Xi Huang, Yuanming Pan, Jin-Xiao Mi

PII: S1387-1811(14)00427-2
DOI: http://dx.doi.org/10.1016/j.micromeso.2014.08.002
Reference: MICMAT 6685

To appear in: Microporous and Mesoporous Materials

Received Date: 23 June 2014


Revised Date: 31 July 2014
Accepted Date: 1 August 2014

Please cite this article as: J-Q. Wang, Y-X. Huang, Y. Pan, J-X. Mi, Hydrothermal synthesis of high purity zeolite
A from natural kaolin without calcination, Microporous and Mesoporous Materials (2014), doi: http://dx.doi.org/
10.1016/j.micromeso.2014.08.002

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.
Hydrothermal synthesis of high purity zeolite A
from natural kaolin without calcination

Jing-Quan Wanga, Ya-Xi Huanga, Yuanming Panb, and Jin-Xiao Mia,*

a
Fujian Provincial Key Laboratory of Advanced Materials, Department of Materials

Science and Engineering, College of Materials, Xiamen University, Xiamen 361005,

Fujian Province, People’s Republic of China

b
Department of Geological Sciences, University of Saskatchewan, 114 Science Place,

Saskatoon, Canada SK S7N 5E2

Abstract

Zeolite A is a commercially and environmentally important microporous material,

which is commonly synthesized from calcined kaolin. One major drawback inherent to

this approach is that the properties (i.e., purity, particle size, morphology, and

cation-exchange capacity) of the final product depend largely on the quality of the raw

kaolin ores. Herein we report on a new hydrothermal route for the synthesis of high

purity zeolite A with controllable particle size and morphology, from natural kaolin

without conventional high temperature calcination. This route starts with hydrothermal

reaction of natural kaolin with NaOH to form hydrous sodium aluminosilicates (i.e.

1
hydroxycancrinite and nepheline hydrate), which are then dissolved in dilute HCl. The

resulting acid solution, after filtration to remove unreacted impurity minerals such as

quartz and muscovite, is then adjusted to pH = 7 to form an amorphous Si and Al gel,

which is used for the hydrothermal synthesis of zeolite A. A comparative study of

zeolite A samples synthesized from this new route and those from the conventional

calcination method reveals that the former have more controllable size and morphology

than the latter, giving rise to higher brightness and enhanced cation-exchange capacity.

Another major advantage of this new method is that the common impurity minerals

such as quartz, muscovite and feldspars in natural kaolin can be removed, giving rise to

high purity zeolite A and possibly expanding this technique to the use of low-grade

kaolin.

Key words: Zeolite A; kaolin; hydrothermal synthesis; NaOH digestion; impurity

removal.

1. Introduction

Zeolites are microporous aluminosilicates widely used as commercial adsorbents

in nuclear fuel reprocessing, [1, 2], catalysts in the petrochemical industry [3-5],

ion-exchange agents for water purification and waste treatments [6, 7], and other

applications [8, 9]. Zeolite A, (Na12[(AlO2)12(SiO2)12]·27H2O), for example, has now

2
replaced phosphates in household detergents, because the latter cause eutrophication of

lakes and rivers by promoting the excessive growth of algae and plankton and adversely

affecting the water quality. Zeolite A possesses the so-called α-cage comprised of 8

cuboctahedra linked by 12 cuboids, which gives rise to excellent ion exchange

capacities without any known adverse impacts on the environment. Silver-loaded

zeolite A, prepared by the ion exchange method, is employed as antimicrobial agent,

with controlled-release function and excellent sterilization [10, 11]. In addition, zeolite

A has been proven to be an efficient ion exchange membrane in vanadium redox flow

batteries [12].

The initial syntheses of zeolite A made use of chemical reagents such as

tetraethylorthosilicate(TEOS), Na2SiO3, NaAlO2 and Al(OH)3 [13]. The principal

chemical components of zeolite A are silica and alumina, which are among the most

abundant and widely available materials in rocks and minerals [14]. It is not surprising

that most current methods for the commercial synthesis of zeolite make use of natural

materials such as kaolinite [15, 16], halloysite [17], palygorskite [18] and diatomite

[19] as well as industrial and agricultural wastes (e.g. coal ash and rice husk) [20, 21].

Among them, the relatively inexpensive and widely available kaolin clay is particularly

attractive, because its principal mineral kaolinite shares the appropriate Si/Al ratio with

zeolite A and has an abundant supply.

The conventional technique for the synthesis of zeolite A from natural kaolin

commences with calcination (i.e. thermal treatment at 700~900 °C) to reduce its

3
chemical-attack resistance and to transform it into an amorphous material (i.e.

metakaolin) [6]. The resulting metakaolin is then used to synthesize zeolites by

hydrothermal reaction in alkaline media [8]. Calcination (also known as “thermal

activation”) not only adds significant financial costs to the synthesis of zeolite A from

kaolin but also has adverse environmental effects (i.e., consumption of large quantities

of fossil fuels and production of large amounts of greenhouse gases)[22, 23].

Another drawback in the conventional synthesis of zeolite from natural kaolin is the

common presence of impurities such as quartz, muscovite and feldspars, which are

inherited from the raw materials [18]. These impurity minerals have higher thermal and

chemical stabilities than kaolinite, making them difficult to eliminate during the

thermal treatment or hydrothermal synthesis [18, 24, 25].

Wang and Tomita introduced a method to remove kaolinite and quartz from

Australian coals by Ca(OH)2/HCl leaching. Kaolinite and quartz in coals were

dissolved after hydrothermal reaction with Ca(OH)2 at 150-300 °C and were converted

to calcium-bearing compounds such as hibschite, omisteinbergite, tobermorite and

jaffeite, which are readily dissolved and leached in HCl solutions [26-28]. These

interesting results inspired the present authors to develop a new hydrothermal route

making use of natural kaolin without calcination, but achieving effective removal of

impurity minerals, for the synthesis of high purity zeolite A.

Microporous materials such as zeolites display attractive optical, electrical,

magnetic, and catalytic properties that often depend strongly on the particle size and

4
shape [29]. Consequently, synthesis of these materials with controllable sizes and

morphologies are highly desirable. Another major objective of this contribution is

therefore the optimization of the new hydrothermal route for the synthesis of high

purity zeolite A with controllable particle size and morphology. This new method –

alkaline pretreatment followed by acid dissolution, has been evaluated by comparison

with the conventional technique using high temperature calcination. Synthetic zeolites

from the new and conventional routes have been characterized by X-ray diffraction

analysis (XRD), Fourier-transform infrared spectroscopy (FTIR), scanning electron

microscopy (SEM), energy-dispersive X-ray spectroscopy (EDX), reflectance, and

cation exchange capacity (CEC) analysis, and are compared and discussed in detail

herein.

2. Experimental methodology

2.1. Kaolin pretreatment

The raw material of natural kaolin used in this study was provided by the LongYan

kaolin Limited Company (Fujian, China). The chemical and mineralogical composition

of the as-received kaolin is given in Table 1. The molar Si to Al ratio of the raw material

is 1.19, well suited for the synthesis of zeolite A. This natural kaolin material was first

hydrothermally treated by NaOH digestion, followed by leaching with dilute HCl.

After removal of the unreacted solid residue, the resulting acid solution was adjusted

by NaOH to precipitate an amorphous Si and Al gel with the proper ratio of Al/Si,

5
which was then used for hydrothermal synthesis of zeolite A.

The kaolin pretreatment involves mainly an alkaline digestion followed by acidic

leaching and can be divided into three steps (Fig. 1).

The first step activates the as-received kaolin raw material with alkaline solutions

of different NaOH concentrations. Two sets of four samples were prepared by mixing 3

g kaolin raw material each with 16 mL of 1, 2, 3 and 4 M aqueous NaOH solutions, via

vigorous stirring. The homogeneous mixtures were then heated at 10ºC/min in a 100

mL closed Teflon-lined stainless steel autoclave, and kept for 3 hours at 190 °C for one

set and 240 °C for the other set. After this step of NaOH digestion, the autoclave was

taken immediately out of the stove and rapidly cooled with tap water. The resulting

materials were then filtered out, rinsed with deionized water, and dried in a desiccator at

90 °C overnight. The solid materials from this alkaline activation were characterized by

X-ray powder diffraction (XRD) analysis (Figs 2a and 3a). Optimization of the

digestion conditions (i.e. hydrothermal-treatment temperature and duration, NaOH

concentrations etc.) was performed, and discussed in the next section.

The second step started with dispersing the solid materials from the first step in

water to form slurries by vigorous stirring and was followed by dissolution in 60 mL of

0.2 M HCl for 0.5 hour under stirring at room temperature. After filtering out the

insoluble impurities, a transparent solution was recovered. The insoluble impurities

were also recovered and weighed for characterization by XRD (Figs 2b and 3b). The

proportion of kaolin dissolution (Table 2) was defined by (1-x)%, where x is the weight

6
ratio of the remaining insoluble impurities to the original kaolin raw material.

The third step took the filtered solution and adjusted it to pH = 7 by using NaOH

while stirring. This step resulted in the precipitation of a white gel, which was filtered

out and characterized by XRD (Fig. 4a1).

2.2. Synthesis of zeolite A

The white gel obtained from the above experimental steps was dissolved in 1.0

M NaOH solution for subsequent hydrothermal crystallization at 90 °C in a sealed

Teflon-lined stainless steel autoclave for 3 days. After hydrothermal crystallization, the

solid product, zeolite A, was filtered out, washed with deionized water, and dried at

90 °C overnight.

For comparison, the same amount of the kaolin raw material was used for the

synthesis of zeolite A by using the conventional route [6]. The kaolin raw material (3 g)

was first calcined at 800 °C for 3 hours. The resulting metakaolin, after pulverizing and

sieving (500 mesh), was mixed with 50 mL of 1 M NaOH solution for hydrothermal

crystallization at 90 °C for 3 days as described above.

2.3. Characterization of zeolite A

The XRD patterns of the kaolin raw material, solids from both steps of alkali

activation and acid dissolution, and synthetic zeolite A were collected on a PANalytical

X'Pert instrument for 2θ’s of 3~65° with a step of 0.016° (2θ) using Cu (Kα) radiation
7
(λ = 0.154184 nm). Fourier transform infrared (FTIR) spectra were recorded in the

range 4000~400 cm−1 on a Nicolet 330 FTIR spectrometer using pellets of 0.5 mg

powder samples mixed with 250 mg KBr. Scanning electron microscopy (SEM) images

and energy dispersive X-ray spectroscopy (EDX) were taken on a LEO-1530 SEM

scanning electron microscope. The samples for SEM imaging were coated with gold.

The compositions of the raw material and synthetic zeolite A were determined by

inductively coupled plasma atomic emission spectrometry (ICP-AES). The kaolin raw

material and synthetic zeolite A were also measured for brightness, which is the

percentage of reflectance at the wavelength of 457 nm relative to a whiteboard (a

standard ceramic panel (85.5%), which has been calibrated by use of BaSO4).

The cation exchange capacity (CEC) was determined by dispersing 0.5 g synthetic

zeolite A in 500 mL of 0.005 mol/L CaCl2 solution for 30 minutes with continuous

stirring (400 rpm) at room temperature. After removal of the sample by filtration, the

Ca2+ concentration in the solution was determined by EDTA titration.

3. Results and discussion

The new route for hydrothermal synthesis of zeolite A has been developed and

optimized on the basis of our systematic experiments that determined factors

controlling the reactivity of kaolin in alkalis and acids. For example, our experiments

show that one key issue during alkaline activation is the ratio between kaolin and

8
NaOH solution, which must be optimized to ensure a complete reaction of kaolinite but

keep the resulting products (i.e. hydroxycancrinite and nepheline hydrate) in the solid

form. The latter allows efficient filtration of the resulting products to remove the NaOH

solution, which otherwise would require the use of excessive HCl in the subsequent

acid dissolution.

3.1 Effects of hydrothermal treatment on kaolin

Powder XRD analyses (Figs 2 and 3) show that the kaolin raw material has been

transformed, partially or almost completely, to hydroxycancrinite

(Na4Al3Si3O12(OH)·H2O, PDF#46-1457) and nepheline hydrate (Na2Al2Si2O8·H2O,

PDF#10-0460) (Figs. 2~3)), after alkali treatment with various NaOH concentrations

(1M, 2M, 3M and 4M) at 190 °C or 240 °C. The solid products from this hydrothermal

treatment vary widely with both the NaOH concentration and temperature.

For example, the solid products from treatments in 1M, 2M and 4M NaOH

solutions at 190 °C contain mainly hydroxycancrinite, but nepheline hydrate

predominates after treatment in the 3M NaOH solution at this temperature (Fig. 2a).

XRD patterns (Fig. 2b) reveal that the solid residue after acidic dissolution

consists of mainly quartz and muscovite, which are the main mineral impurities in

kaolin (Table 1) and are more stable than kaolinite in the alkaline and acidic media.

Quartz and muscovite have been detected in all XRD patterns of samples treated in all

NaOH concentrations, whereas kaolinite is present only in the sample treated with 1 M

9
NaOH. Quartz and muscovite are only partially dissolved and altered to

hydroxycancrinite and nepheline hydrate in alkaline solutions, whereas kaolinite was

completely reacted in alkaline solutions except for the case of 1 M NaOH (Fig. 2b1).

Interestingly, hydroxycancrinite and nepheline hydrate are not detected in the XRD

patterns of the solid residue after acid dissolution (Fig. 2b), indicating their complete

dissolution in 0.2 M HCl.

At 240 °C, the contents of quartz and muscovite impurities are clearly decreased

with the increase of the NaOH concentration (Fig. 3a). For example, neither muscovite

nor quartz or kaolinite is detectable in the XRD pattern of the sample treated with 4 M

NaOH (Fig. 3a4). Also, virtually no solid residue is present after acid dissolution,

suggesting that kaolinite as well as the impurity minerals have been converted to

nepheline hydrate and hydroxycancrinite after treatment with 4 M NaOH at 240 °C.

Table 2 shows that the proportion of kaolin dissolution varies widely with the

condition of hydrothermal treatments. The kaolin dissolution increases with the

increase of temperature and the NaOH concentration. For example, the proportions of

kaolin dissolution with 2 M NaOH at 190 °C and 240 °C are 69.7% and 82%,

respectively, and no kaolinite was detected in the XRD patterns of the solid residue (Fig.

2b2 and Fig. 3b2). These results collectively show that the optimal condition for a

complete dissolution is the use of 4 M NaOH at 240 °C for 3 hours.

10
3.2 Comparison of products from the two synthesis methods

The final products from the two synthesis methods described above have similar

FTIR spectra (Fig.4b), which are consistent with zeolite A. The bands at 466 cm−1 and

476 cm−1 in Figures 4b2 and 4b1 are attributable to the internal TO4 (T = Si, Al)

tetrahedral bending modes, whereas the bands at 556 and 553 cm−1 are assigned to the

vibration modes of the double rings in zeolite A. The intense bands at 1003 and 996

cm−1 are attributable to the T-O asymmetric stretching modes [30-32] . The broad

bands at 3445 and 3446 cm−1, along with the weak bands at 1641 and 1639 cm−1, arise

from the water molecules and the OH groups in zeolite A. Their XRD patterns (Fig. 5a)

are also broadly similar and further confirm successful synthesis of zeolite A. These

XRD patterns show that the final products from the new method are almost pure zeolite

A, except for the presence of minor zeolite P identified by the weak reflection at 2θ =

28.1° (Fig. 5a1). The synthetic product from the conventional calcination technique, on

the other hand, contains additional reflections at 19.76° and 26.64°, which are the

characteristic peaks of muscovite and quartz, respectively (Fig. 5a2). It is interesting to

note that the weak reflection of zeolite P is also present in the XRD pattern of the

product from the conventional technique (Fig. 5a2). The presence of zeolite P as a

minor impurity phase in the final products from both techniques has also been

confirmed by SEM imaging, where zeolite A has a diagnostic cube morphology

different from zeolite P in small aggregated spheres (Fig. 5a). These results

demonstrate that zeolite A from the new method has a higher purity than its counterpart

from the conventional technique.


11
The quartz and muscovite impurities in the final product from the conventional

technique can be readily attributed to their resistance to calcination, which has been

confirmed by XRD analyses of the intermediate products. Specifically, XRD patterns

of kaolin calcined at 800 °C for 3 hours show that kaolinite has been completely

transformed to metakaolin (MK), but the characteristic peaks of muscovite and quartz

remain (Fig. 4a2). In contrast, the white gel from the new method after alkaline

activation and acid dissolution is completely amorphous (Fig. 4a1), without any

evidence of kaolinite or impurity minerals.

The identities of the impurity minerals in the product from the conventional

technique have been further evaluated by EDX analysis (Fig. 5b). For example, the

cubic particles (spot 1) containing Na, Al, Si and O are consistent with zeolite A. The

flat particles (spot 2) contain additional K, indicative of muscovite

(KAl2(Si3AlO10)(OH)2) (Fig. 5b).

Figure 6 compares the sizes and morphologies of the products synthesized from

the conventional calcination technique and the new hydrothermal route. The average

particle size (about 3 µm) of the products obtained from the new route (Fig. 6c & d),

is smaller than that of samples (>5 µm) from the conventional method (Fig. 6a & b).

Interestingly, the particles in Fig. 6a & b from the conventional method are

xenomorphic and irregular in shape, whereas those in Fig. 6c & d from the new route

are uniformly euhedral cubes. Therefore, zeolite A particles obtained from the

conventional method are polycrystalline with irregular shapes (Fig. 6a & b), whereas

12
those from the new route are well-developed cubic monocrystallites (Fig. 6c & d).

This difference may be attributed to different nucleation and growth mechanisms of

the two methods [33]. In the new method, nucleation might have taken place

uniformly in the amorphous silica-alumina gel, yielding the monocrystallites [34]. In

the conventional method, the nucleation and crystallization processes of zeolites

probably initiated on metakaolinite precursors [33], which are expected to produce

polycrystallites owing to preferential growth. The polycrystalline material from the

conventional calcination technique is expected to possess more variable surface areas

and defects than monocrystallites, resulting in the blockage of the 3D interconnected

pores and consequently adversely affecting the ion exchange and adsorption capacity

of zeolites (see below).

Moreover, zeolite A synthesized from the same kaolin raw material but using

different methods differs significantly in the Si/Al ratio, brightness and cation exchange

capacity (CEC). The measured Si/Al ratios of zeolite A obtained from the new method

and the conventional technique are 1.33 and 1.80, respectively. Interestingly, their

corresponding FTIR bands at 996 and 1003 cm−1 are consistent with the expected

framework vibrational frequencies of zeolite A, which are known to vary with the

Si/Al ratio [32]. Zeolite A from the new method has a measured brightness of 68.7 %,

higher than that (65.9%) of its counterpart from the conventional technique. The

improved brightness from the new method may be attributed to partial removal of trace

impurities such as Fe2O3 and MnO by filtration, which are known to cause reduced

13
brightness [25, 35]. The CEC of zeolite A from the new method is 299 mg CaCO3/g,

which is significantly higher than that (244 mg CaCO3/g) of its counterpart from the

conventional technique.

It is well known that the properties (i.e., purity, particle size and morphology) of

the final product from the conventional calcination technique depend strongly on the

quality of the raw kaolin ores. Although the supply of kaolin is virtually unlimited,

super quality kaolin ores have become increasingly rare and expensive. Proper

utilization of low-quality kaolin ores is therefore becoming more and more important.

Industrial techniques for the removal of adverse impurities such as Fe2O3 from

low-grade kaolin ores have long been developed and widely applied [36, 37]. These

techniques can be broadly divided into two types: chemical or physical methods. The

former is mainly based on the deoxidization of sodium dithionite (Na2S2O4) with

addition of H2SO4 [36], whereas the latter usually makes use of different magnetic

properties of minerals for separation [37]. These established methods, however, have

little effects on common silicate impurities such as quartz, mica and illite in low-grade

kaolin ores. The new hydrothermal route reported herein can not only effectively

remove these silicate impurities but also partially convert them to useful ingredients

(i.e. Al2O3 and SiO2) for the synthesis of zeolite A. Therefore, the new hydrothermal

route developed in this study has the potential for a wide utilization of low-grade

kaolin ores. Utilization of low-grade kaolin ores that are widely available and low in

cost is obviously advantageous for the mass production of zeolite A. Indeed, one main

14
problem in the production of all types of synthetic zeolites is the availability and cost of

raw materials.

Finally, many other industrially and environmentally important zeolites, such as

barrerite, chabazite, clinoptilolite, faujasite, heulandite, mordenite, natrolite, and

stilbite, have Si, Al and Na as their principal components. The success of our new

hydrothermal route also opens the door for supplying high-quality sodium

aluminosilicate gels for the synthesis of other zeolites (e.g. X, Y and P).

4. Conclusions

We have successfully developed and optimized a new hydrothermal route to

synthesize high-purity zeolite A from natural kaolin. The new route consists of four

steps, with the optimal conditions including: (1) alkaline activation with 4 M NaOH at

240 °C for 3 hours; (2) acid dissolution in 0.2 M HCl for 0.5 hour; (3) adjusting the acid

solution by adding Na(OH) to pH=7 to form a gel; and (4) the gel dissolved in 1 M

NaOH for hydrothermal synthesis at 90 °C for 3 days. The primary advantage of this

route is that the impurity minerals such as muscovite and quartz in the kaolin raw

material can be eliminated or transformed to the chemical components (i.e., silica and

alumina) of the target products. This elimination of the common impurity minerals in

the raw material not only gives rise to high purity zeolite A but also allows the use of

low-grade kaolin, and possibly opening up other applications of low-grade kaolin

15
materials as well. In addition, the zeolite A product from this new route has more

uniform particle sizes and morphologies, higher brightness, and enhanced

cation-exchange capacity than its counterpart from the conventional calcination

technique.

5. Acknowledgements

We thank the National Natural Science Foundation of China (Nos. 40972035,

21233004 & 21201144), the Fundamental Research Funds for the Central Universities

(2013121020), and Natural Science and Engineering Research Council of Canada for

financial support.

6. References

[1] A. Dyer, D. Keir, Zeolites, 4 (1984) 215-217.


[2] H. Yeritsyan, A. Sahakyan, V. Harutyunyan, S. Nikoghosyan, E. Hakhverdyan, N. Grigoryan, A.
Hovhannisyan, V. Atoyan, Y. Keheyan, C. Rhodes, Sci. Rep., 3 (2013) 2900.
[3] A. Corma, Chem. Rev., 97 (1997) 2373-2420.
[4] P. Dornath, W. Fan, Micropor. Mesopor. Mat, 191 (2014) 10-17.
[5] M.C. Silaghi, C. Chizallet, P. Raybaud, Micropor. Mesopor. Mat, 191 (2014) 82-96.
[6] E. Costa, A. De Lucas, M.A. Uguina, J.C. Ruiz, Ind. Eng. Chem. Res., 27 (1988) 1291-1296.
[7] Z.T. Xue, Z.L. Li, J.H. Ma, X. Bai, Y.H. Kang, W.M. Hao, R.F. Li, Desalination., 341 (2014) 10-18.
[8] D.W. Breck, Zeolite molecular sieves: Structure, Chemistry, and Use, John Wiley & Sons Inc, New
York, 1984.
[9] S.K. Hoffmann, J. Goslar, S. Lijewski, K. Tadyszak, A. Zalewska, A. Jankowska, P. Florczak, S. Kowalak,
Micropor. Mesopor. Mat, 186 (2014) 57-64.
[10] Y.Z. Zhou, Y.Q. Deng, P. He, F.Q. Dong, Y.J. Xia, Y. He, Rsc Advances, 4 (2014) 5283-5288.
[11] A. Mayoral, T. Carey, P.A. Anderson, I. Diaz, Micropor. Mesopor. Mat, 166 (2013) 117-122.
16
[12] Z. Xu, I. Michos, X.R. Wang, R.D. Yang, X.H. Gu, J.H. Dong, Chem. Commun., 50 (2014) 2416-2419.
[13] M.E. Davis, R.F. Lobo, Chem. Mater., 4 (1992) 756-768.
[14] R.M. Barrer, R. Beaumont, C. Collela, J. Chem. Soc., Dalton Trans., (1974) 934-941.
[15] M. Alkan, Ç. Hopa, Z. Yilmaz, H. Güler, Micropor. Mesopor. Mat, 86 (2005) 176-184.
[16] K. Shams, H. Ahi, Micropor. Mesopor. Mat, 180 (2013) 61-70.
[17] Y.F. Zhao, B. Zhang, X. Zhang, J.H. Wang, J.D. Liu, R.F. Chen, J. Hazard. Mater., 178 (2010) 658-664.
[18] J.L. Jiang, L.D. Feng, X. Gu, Y.H. Qian, Y.X. Gu, C.S. Duanmu, Appl. Clay. Sci., 55 (2012) 108-113.
[19] Y.C. Du, S.L. Shi, H.X. Dai, Particuology, 9 (2011) 174-178.
[20] H. Katsuki, S. Komarneni, J. Solid State Chem., 182 (2009) 1749-1753.
[21] X. Querol, F. Plana, A. Alastuey, A. Lopez-Soler, Fuel, 76 (1997) 793-799.
[22] C.R. Melo, H.G. Riella, N.C. Kuhnen, E. Angioletto, A.R. Melo, A.M. Bernardin, M.R. da Rocha, L. da
Silva, Mat. Sci. Eng. B, 177 (2012) 345-349.
[23] C.A. Ríos, C.D. Williams, M.A. Fullen, Appl. Clay. Sci., 42 (2009) 446-454.
[24] D. Akolekar, A. Chaffee, R.F. Howe, Zeolites, 19 (1997) 359-365.
[25] H.H. Murray, Clay. Miner., 34 (1999) 39-49.
[26] J. Wang, A. Tomita, Ind. Eng. Chem. Res., 36 (1997) 5258-5264.
[27] J. Wang, A. Tomita, Ind. Eng. Chem. Res., 36 (1997) 1464-1469.
[28] J. Wang, A. Tomita, Fuel, 77 (1998) 1747-1753.
[29] M. Ansari, A. Aroujalian, A. Raisi, B. Dabir, M. Fathizadeh, Adv. Powder Technol., 25 (2014)
722-727.
[30] M.L. Granizo, M.T. Blanco-Varela, S. Martinez-Ramirez, J. Mater. Sci., 42 (2007) 2934-2943.
[31] R. Szostak, Handbook of molecular sieves, Van Nostrand Reinhold, New York, 1992.
[32] P.K. Dutta, B. Del Barco, J. Phys. Chem., 92 (1988) 354-357.
[33] J. Rocha, J. Klinowski, J.M. Adams, J Chem. Soc. Faraday Trans., 87 (1991) 3091-3097.
[34] L. Itani, Y. Liu, W.P. Zhang, K.N. Bozhilov, L. Delmotte, V. Valtchev, J. Am. Chem. Soc., 131 (2009)
10127-10139.
[35] S. Chandrasekhar, P. Raghavan, G. Sebastian, A.D. Damodaran, Appl. Clay. Sci., 12 (1997) 221-231.
[36] W.Q. Gong, X.H. Wu, J. Wuhan. Univ. Technol, 12 (1997) 47-57.
[37] M.F. Wang, Z. Zhu, F.P. Ning, H. Yang, G.Q. Zhang, K.X. Wang, J. Supercond. Nov. Magn., 27 (2014)
1397-1401.

17
Table 1. Chemical and mineralogical composition of the raw kaolin material
Oxides SiO2 Al2 O3 K2O Fe2O3 MgO CaO Na2O TiO2 H2O Total
wt% 49.64 35.29 2.72 0.27 0.22 0.20 0.11 0.03 11.52 100
Minerals Kaolinite Muscovite Quartz Feldspars Others Total
wt% 69.7 18.7 6.8 3.2 1.6 100

Table 2.Kaolin dissolution under various hydrothermal treatments


Hydrothermal alkali digestion (for 3 h) Acid leaching (0.2 M HCl at 25°C for 0.5 h)
Temp. NaOH Products* Residues Kaolin dissolution
190°C 1M C, M, Q, K M, Q, K, F 56.2 %
190°C 2M C, M, Q, N, P M, Q 73.0 %
190°C 3M C, M, Q, N M, Q 77.4 %
190°C 4M C, M, Q, N, S M, Q 78.1 %
240°C 1M C, M, Q, K M, Q, K, 60.8 %
240°C 2M C, M, Q, N M, Q 82.0 %
240°C 3M C, N M, Q 88.5 %
240°C 4M C, N — 99.9 %
*
Abbreviation: K denotes kaolinite; M: muscovite; Q: quartz; F: feldspars; N:
nepheline hydrate; C: hydroxycancrinite; S: sodalite; P: zeolite P.

Figure captions:

Figure 1. Flow chart summarizing the synthesis steps of zeolite A. R1 for the new route
of hydrothermal reaction; R2 for the conventional high-temperature calcination route.

Figure 2. Comparison of XRD patterns of kaolin activated at 190 °C with 1 M, 2 M, 3


M and 4 M NaOH solutions (Left a1, a2, a3, a4), and the corresponding insoluble solid
residues obtained after HCl dissolution (Right b1, b2, b3 and b4). The vertical bars at
the bottom mark the Brag positions of hydroxycancrinite(PDF#46-1457; denoted as
C), nepheline hydrate(PDF#10-0460; N), muscovite (PDF#07-0025; M), and kaolinite
(PDF#14-0164; K). Abbreviations are same as those in Table 2.

Figure 3. Comparison of XRD patterns of kaolin activated at 240 °C with 1 M, 2 M, 3


M and 4 M NaOH solutions (Left a1, a2, a3, a4), and the corresponding insoluble solid
residues obtained after HCl dissolution (Right b1, b2, b3 and b4). The vertical bars at

18
the bottom mark the Brag positions of hydroxycancrinite(PDF#46-1457; denoted as
C), nepheline hydrate(PDF#10-0460; N), muscovite (PDF#07-0025; M), and kaolinite
(PDF#14-0164; K). Abbreviations are same as those in Table 2.

Figure 4 Left (a): comparison of XRD patterns of raw kaolin (a3, upper), matakaolin
(a2, middle) obtained from the conventional technique, and amorphous Si and Al gel
(a1, bottom) obtained from the new method. Right (b): comparison of FTIR spectra of
zeolite A from the novel route (b1, bottom) and conversional technique (b2, upper).

Figure 5 Comparison of XRD patterns and SEM images of the products obtained from
the conversional calcination technique (left upper, a2) and the new hydrothermal (left
bottom, a1). The chemical composition of the particle from a conversional technique
further identified by EDX (right). Spot 1 for zeolite A and spot 2 for muscovite.
Designation M: muscovite; Q: quartz; P: zeolite P.

Figure 6 Comparison of SEM images of zeolite A obtained from the conversional


calcination technique (a, b) and the new hydrothermal route (c, d). Note that
morphologies in (a) and (b) are xenomorphic, whereas those in (c) and (d) are
euhedral cubes.

19
Fig_1_color_online
Figure_2
Figure_3
Figure_4
Figure_5
Fig_6
Highlights (<85 characters)
(1) A new hydrothermal route to synthesize high-purity zeolite A from natural kaolin;

(2) Optimization to ensure complete kaolinite dissolution and removal of impurities;

(3) Zeolite A of controllable size and morphology, and enhanced cation-exchange

capacity;

(4) Potential extensions to use of low-grade kaolin and synthesis of other zeolites.

20
Graphical abstract legend:

A new hydrothermal route has been developed for the synthesis of high purity

zeolite A with controllable particle size and morphology, from natural kaolin without

high temperature calcination.

21

You might also like