Next Article in Journal
Hydration Kinetics of Portland Cement–Silica Fume Binary System at Low Temperature
Previous Article in Journal
Evaluation of Thermally Treated Calotropis Procera Fiber for the Removal of Crude Oil on the Water Surface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of X-Zeolite from Waste Basalt Powder and its Influencing Factors and Synthesis Mechanism

1
School of Civil Engineering, University of South China, Hengyang 421001, China
2
Key Laboratory of Special High Performance Concrete, University of South China, Hengyang 421001, China
3
School of Chemistry and Chemical Engineering, University of South China, Hengyang 421001, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(23), 3895; https://doi.org/10.3390/ma12233895
Submission received: 26 October 2019 / Revised: 12 November 2019 / Accepted: 21 November 2019 / Published: 26 November 2019
(This article belongs to the Section Porous Materials)

Abstract

:
Traditional hydrothermal method (TH) and alkali fusion-assisted hydrothermal method (AFH) were evaluated for the preparation of zeolites from waste basalt powder by using NaOH as the activation reagent in this study. The synthesized products were characterized by BET, XRD, FTIR and SEM. The effects of acid treatment, alkali/basalt ratio, calcination temperature and crystallization temperature on the synthesis process were studied. The results showed that AFH successfully synthesized zeolite X with higher crystallinity and no zeolite was formed by TH. The specific surface area of synthetic zeolite X was 486.46 m2·g−1, which was much larger than that of original basalt powder (12.12 m2·g−1). Acid treatment and calcination temperature had no effect on zeolite types, but acid treatment improved the yield and quality of zeolite. Alkali/basalt ratio and crystallization temperature not only affected the crystallinity of synthesized zeolites but also affected its type. The optimum synthesis condition of zeolite X are as follows: acid treatment of 5 wt% HCl solution, NaOH/basalt ratio of 1:1, a calcination temperature of 650 °C and crystallization temperature of 120 °C. The work shows that basalt can be used as a raw material to prepare zeolite.

1. Introduction

Zeolites, three-dimensional tetrahedral hydrated aluminosilicate minerals with meso and microporous structures, are widely used as catalysts in separation and refinery industries [1,2,3], feed additives in animal husbandry [4,5], adsorbents in wastewater treatment for the removal of heavy metal cations, anions and dyes [6,7,8]. However, the cost of industrial synthetic zeolites from chemical sources is pretty high, which limits its application to a large extent. Therefore, a lot of attentions have been paid on searching for cheap raw materials for zeolites preparation in recent years. Thirty years ago, Barth-Wirsching and Ulrike [9] achieved the zeolite conversion in the laboratory by simulating the formation of natural zeolites, laying a solid foundation for further study.
So far, various geomaterials and industrial wastes have been used as starting materials for zeolites preparation, such as: Natural kaolin [10], coal gangue [11], coal fly ash [12], bentonite [13], shale rock [14], paper sludge ash [15], bagasse fly ash [16], halloysite [17], waste porcelain [18], lithium slag [19] and so on. With the exploration of raw materials, the synthesis methods of zeolites have been continuously developed. The common methods are traditional hydrothermal method [19], alkali fusion method [20], alkali fusion-assisted hydrothermal method [21], two-step method [22], and sonochemical method [23]. Moreover, in terms of the zeolite synthesis mechanism, there are three relatively developed types, namely solid-phase transition mechanism, liquid-phase transition mechanism and solid-liquid phase transition mechanism.
Coal fly ash (CFA) and other materials are chosen as the starting materials for zeolites synthesis given the high content of Si or/and Al, and given the high content of reactive phases, such as amorphous aluminosilicate glasses. In terms of the standard features, basalt may serve as another ideal raw material. It is the similarity in the chemical composition of basalt and CFA that leads to the testing of basalt based zeolites growth. As well, a large amount of basalt reserves in China make basalt an inexpensive material. Furthermore, the synthesis of zeolites from basalt may enhance its potential value and extend its serviceable scope and reduce the cost of traditional industrial synthetic zeolites. For now, there are few works in literature concerning the synthesis of zeolites from basalt.
The present study centers on the synthesis of zeolites from wasted basalt powder by using TH and AFH, the main objectives being: (a) To obtain an effective synthesis method; (b) to illuminate the activation conditions for maximum synthesis efficiency; (c) to analyze the effects of different synthesis conditions; (d) to reveal the synthesis mechanism of basalt based zeolites.

2. Materials and Methods

2.1. Materials

Water, used in the whole experiment, is deionized. Sodium hydroxide (NaOH) and hydrochloric acid (HCl) mentioned below are analytically pure. For CFA, NaOH is more useful to stimulate its activity than KOH and Na2CO3, hence NaOH is used as the activation reagent in this study. Besides, in order to mix with basalt powder easily, NaOH adopted is granular.
Waste basalt powder was gathered in batches from a basalt mine located in Hunan Province, China. All the collected basalt was mixed well together and ground into fine powders with particle sizes of less than 50 µm (300 mesh sieve).

2.2. Synthesis Process

2.2.1. Pretreatment

To study the transition process and mechanism in an even better fashion, some pretreatments of raw basalt powder are designed as follows.
Sample A: Raw basalt powder (particle size ≤50 µm), without treatment. Sample B: Take some sample A into a crucible, annealed in a muffle furnace at 650 °C with a temperature increase rate of 11.5 °C·min−1 for 3 h. Mark the annealed solid as sample B. Sample C: Wash some sample A with prepared 5 wt% HCl solution in a sealed beaker at room temperature, followed by vigorous magnetic agitation with a rate of 350 r·min−1 for 4 h, in which the liquid/solid ratio is 3:1 (30 mL/10g). After that, separate the solid by centrifugation and wash it with deionzied water in excess. Then dry the washed solid in an electrothermal drying oven at 80 °C for 10 h. Mark the dried product as sample C. Sample D: Treat some sample C with the same method and procedure described in sample B. Then we get sample D. Sample E: Mix 10 g of sample C and 10 g of NaOH uniformly, and treat the mixture with the same method and procedure described in sample B. Grind the annealed lump into powders, namely sample E.

2.2.2. Traditional Hydrothermal Method

This method comprises two major stages: aging and crystallization. First, 10 g of raw material (sample A, B, D) was mixed with 50 mL of 5 M NaOH solution in a beaker. Then the beaker, sealed with preservative film, was set in a water bath for aging at 35 °C with sustained magnetic stirring with a rate of 350 r·min−1 for 10 h. Second, transfer the aged mixture into a closed steel vessel lined with PTFE, crystallizing at 120 °C for 12 h. The solid was recovered, washed and dried in the oven at 80 °C.

2.2.3. Alkali Fusion-Assisted Hydrothermal Method

First, 10g of starting material (sample A, sample C), mixed homogeneously with a certain amount of NaOH pellets (5 g, 10 g, 15 g, 20 g) in a crucible, was calcined for alkali fusion in the muffle furnace at a certain temperature (550–750 °C) for 3 h. Then, grind the alkalized lump into powder with a mortar. Blend the powder with 50 mL deionized water for maturing at constant shaking in the 35 °C water bath for 10 h. After that, transfer the matured mixture into a closed steel vessel lined with PTFE for crystallization at a certain temperature (25 °C, 90–150 °C) for 12 h. Last, rinse the crystallographic product with deionized water for a few repetitions. During washing, separate the supernatant precipitate from the lower precipitate. Then dry the washed supernatant precipitate.

2.3. Characterization

The bulk chemical compositions of raw basalt sample were determined by X-Ray Fluorescence (XRF, Model: Axios PW4400, Holland). The mineralogical composition of starting materials, intermediate product and synthetic products were analyzed by X-ray diffraction (XRD, Model: Bruker D8, Germany) with diffraction angle (2θ) ranging from 5° to 90°. The minerals of samples were identified by matching actual spacing of an unknown mineral. Surface morphology analysis of basalt and synthetic products was performed by scanning electron microscope (SEM, Model: JSM-7500F, Japan). The FTIR spectrum was achieved by means of a FTIR spectrometer (FTIR, Model: Nicolet-460, USA) ranging from 400 cm−1 to 4000 cm−1. In addition, N2 adsorption-desorption was undertaken at 77 K by a volumetric absorption analyzer (BET, Model: ASAP 2020, USA) and the surface area was measured by the BET method.

3. Results and Discussion

3.1. Properties of Raw Basalt Powder

The bulk chemical composition of raw basalt powder was determined by XRF and given in Table 1. As shown in Table 1, oxides of Si and Al account for about 47.89% and 18.17%, respectively, which act as the fundamental sources of Si and Al needed for zeolites formation. In addition to Al and Si, Fe is the third major element and has an oxide content of 14.67%.
As can be seen from the X-ray diffractogram in Figure 1a, basalt sample is an amorphous phase based on an aluminosilicate glass with minor amounts of crystalline phase substances, mainly labradorite (JCPDS NO. 43-1368) and andesine (JCPDS NO.79-1149). Furthermore, N2 adsorption-desorption isotherm of sample A is given in Figure 1b, indicating that sample A has a dense structure with few pores and channels. The specific surface area of raw basalt powder was measured by BET to be about 12.12 m2·g−1, as shown in Table 2. In addition, some pore-structure parameters are also given in Table 2.
The surface morphology of basalt was characterized by SEM, as shown in Figure 1c,d. As can be seen, the basalt particles are irregular clumps with dense microstructures. Most particles have a relatively smooth surface, and there are no apparent micropores and through holes on the surface, which is consistent with the result from N2 adsorption-desorption isotherm.

3.2. Comparison of the Two Methods

3.2.1. Traditional Hydrothermal Method

Sample A, B and D were adopted as starting materials by using this method. The results showed that synthesized products of sample A, B and D didn’t show any changes in appearance, still in powder form with the same color as before. All the synthetic products were analyzed by XRD and the diffractograms are shown in Figure 2. As can be seen from Figure 2, as the reaction went on, the peak intensity of the crystal phases in sample A gradually declined but did not disappear. Noteworthily, no new peaks were showing. The crystalline phases of all the synthetic products were the same as sample A, without zeolite crystals. No zeolites were synthesized. The results turned out that TH is not an effective way to prepare zeolites from basalt powder.
Nevertheless, this method was successfully performed to form zeolites from CFA by Doyle A M et al. [24]. That it did not work well on basalt is possibly attributed to the quite different surface morphologies of CFA and basalt. The CFA particles are spherical, but basalt particles are mainly irregular lumps with dense microstructures, which means that basalt has a lower contact area for reaction with OH than CFA. On the other hand, the compact inner structure of basalt blocks the spreading pathway for OH, which may weaken the dissolution of Si oxides and Al oxides.
In addition, as can be seen from Figure 2, after the reaction the peak intensity of the crystalline phases decreases, which signifies that some labradorite and andesine were digested during synthesis process. From the differences of the descending extent, we may conclude that calcination and acid treatment activate reactivity of basalt to some extent, concurring with the conclusion on CFA [24].

3.2.2. Alkali Fusion-Assisted Hydrothermal Method

In this method, a mixture of NaOH and starting material (sample A and C) was heated in high temperature (550–750 °C) for alkali fusion before hydrothermal treatment. The XRD pattern of synthesized product is shown in Figure 3. As can be seen in Figure 3, following the synthesis course, the peaks of crystalline minerals in basalt decline and disappear; new crystals occur and grow. The characteristic diffraction peaks of the sample were stable without obvious impurity peak, and the peak width is narrow and sharp, indicating that the product has high purity, complete structure and single crystal phase. In addition, there are seven peaks with strong intensity, at 2θ of 6.3°, 10.1°, 11.9°, 15.6°, 23.5°, 26.8° and 31.1°, respectively. The diffraction peaks are identical with Na2Al2Si2.5O6.2H2O (zeolite X, JCPDS PDF 38-0237) [25]. Zeolite X was successfully synthesized by this method. The experiment result illustrated that AFH is an effective way to synthesize zeolites from basalt powder and alkali fusion is more effective to dissolve the aluminosilicate minerals than hydrothermal treatment, which concurs with the synthesis of kaolin zeolites [26].
Essentially, the fusion of the NaOH and basalt powder mixture facilitates the formation of highly active Na-silicates and Na-aluminates, which are readily dissolvable in aqueous solution and enhance zeolite formation significantly. The higher concentration of aluminosilicates, Na-silicates and Na-aluminates in the reaction system can precipitate zeolites more easily.

3.2.3. Characterization of Zeolite X

Besides XRD pattern above, to study the physicochemical properties of synthetic zeolite X under optimum condition, further characterizations were carried out by SEM, BET and FTIR.
The surface morphological structure of the synthesized product was determined by SEM, as shown in Figure 4. As clearly shown in Figure 4, the crystallites occur and grow with different sizes based on the rotted basalt particle. Lots of ZX crystallites cluster together surrounding the basalt particles, like blossoming flowers. There are only a few single crystals. That is possible because the crystalsare fed and grow in a relatively calm environment without stirring. The crystallites, are mainly octahedral-shaped, accompanied by some littery crystals in atactic shape, which is in accord with the morphology of faujasite (zeolite X is a typical type of faujasite). The morphological analysis of SEM images is in accord with the result of XRD above.
FTIR spectrum of synthesized zeolite X is shown in Figure 5. The single strong adsorption band at 3470.94 cm−1 is attributed to OH stretching of water molecules in the zeolite caves. The band at 1639.94 cm−1 ascribed to H2O deformation mode because of incomplete dehydration of the synthetic zeolite, indicating that there is free water present in the zeolite structure. Moreover, the typical bands, attributed to asymmetric stretch mode (973.24 cm−1), symmetric stretch mode (666.08 cm−1), double six-member rings (D6R, 561.21 cm−1) and a bending mode of the T-O bond (460.11 cm−1), are observed (where T is Al or Si). The spectra data is consent with lithium slag based zeolite X [27].
N2 adsorption-desorption isotherm of zeolite X, as shown in Figure 6, was undertaken to calculate the BET surface area, pore volume and pore size and the result is given in Table 3. From Figure 6, zeolite X is structured with micropores, quite different from basalt shown in Figure 1b. The specific surface area of synthetic zeolite X was measured by BET to be 486.46 m2·g−1, lower than pure zeolite X of 720 m2·g−1 [28], but much larger than basalt sample of 12.12 m2·g−1.

3.3. Effects of Different Synthesis Conditions

3.3.1. Acid Treatment

As shown in Table 1, the basalt sample is a Fe-rich type, with Fe oxides accounting for about 14.67%. According to the related literature [29], iron oxides in the CFA is known to be undesirable for zeolites formation. So acid pretreatment was designed to remove Fe2O3 in the raw basalt powder. Table 4 shows three main components of basalt powder obtained after acid treatment by 5 wt% HCl solution. It indicates a reduction of about 6.5% of iron components and increases of about 11.76% and 6.18% in silicon and aluminum compounds, respectively. After acid washing, the sample color turns light grey from dark grey, which may be explained by the reduction of Fe oxides.
Sample A and sample C were both adopted under the same synthesis condition to study the effects of acid treatment. From Figure 7, zeolites were generated from both sample A and C, with the same single crystalline phase (ZX). However, from the intensity of the peaks, the zeolite product of sample C has a much higher degree of crystallinity than that of sample A. The removal of Fe by acid washing may explain the improvement of crystallinity. In addition, after acid treatment, the yield increases (2.375 g vs. 1.420 g from 10 g of starting material), and the whiteness and fineness of ZX enhances, which is consistent with the prior study of CFA zeolites [29]. Accordingly, acid treatment enhances the yield and quality of synthesized zeolites but does not affect zeolite types.
Acid treatment may also leach a certain amount of Al2O3, which is located in the outer surface of basalt particles. Thus, it relatively improves the SiO2 content of the reaction system and the Si/Al ratio may increase to some extent. However, in this study, the synthetic crystalline phase didn’t show any changes, which may be explained in terms of the process used. In addition, previous works show that faujasite could be formed with Si/Al ratio ranging from 2 to 5 [30].
Since acid treatment is in favor of zeolite formation, sample C is used in the following experiments.

3.3.2. Alkali/Basalt Ratio

Alkali fusion is a general approach to decompose materials rich in silicon or/and aluminum. NaOH present in the reaction system acts as an activator for the formation of soluble aluminate and silicate salts, which are the sources of Si and Al for zeolite synthesis. A series of experiments were undertaken to analyze the effects of NaOH/basalt ratio on zeolite formation and the products were evaluated by XRD as shown in Figure 8. The results showed that there was only a little zeolite (zeolite X) yielded when the alkali/basalt ratio is lower than 0.5:1. However, what is noteworthy is that most of the prior crystalline phases in basalt were melted into amorphous silicates and aluminates compounds, as shown in Figure 8. AFH is more lethal to the crystalline phases and amorphous component in basalt than TH, which verifies the conclusion in comparison of the two methods. When the ratio is up to 1:1, the ZX crystal sprang up like mushrooms with a high degree of crystallinity, with prior crystal phase disappearing. With the NaOH/basalt ratio increasing, the intensity of the ZX reflections decreases without other crystal phases showing up. Following the disappearance of ZX peak, zeolite hydroxylsodalite (HS, JCPDS NO. 41-0009) occurred when the ratio is around 2.0:1. Furthermore, when the ratio is above 2:1, the annealed lump becomes very tough, as well as the annealed product of a pasty mixture of basalt and NaOH, which is very hard to be ground into a powder and not used for further study.
From the disappearance of labradorite and anorthite peaks, it is deduced that both labradorite and anorthite in the basalt powder have reacted with NaOH. Alkali fusion is very effective in extracting silicon and aluminum species in basalt powder, by which labradorite and andesine in basalt were melted in 3 h. However, the fusion of minerals in basalt lies a foundation for zeolites synthesis. Alkalinity and concentration of Na+ play an important role in the synthesis course of zeolites. Low alkalinity (alkali/basalt ratio of 0.5:1) cannot provide the appropriate condition for rearranging Si4+, Al3+ and Na+ into zeolites crystal. Na+ is known as the stabilizer of the sub-building units of zeolite frameworks [30]. Therefore, NaOH/basalt ratio plays a crucial role in the formation of the zeolites, which affects not only the degree of zeolite crystallinity but also the synthesized zeolite types.

3.3.3. Calcination Temperature

Generally, a higher calcination temperature could enhance the amount of Si4+ and Al3+ extracted from basalt powder, as well as the reaction efficiency. However, a previous study on formation of CFA zeolites indicated that too high calcination temperature may go against zeolite formation [30], because the Si and Al fusion compound may convert into other crystalline phases instead of zeolite crystal. So, three moderate temperature levels are set in this study, and the result is given in Figure 9.
As demonstrated in Figure 9, ZX was formed under different calcination temperatures, but with different degrees of crystallinity. When the temperature is lower than 650 °C, the peak intensity increases with the increase of calcination temperature, and then descends over 650 °C without the crystal phases changing. Furthermore, the yield increases gradually as the calcination temperature increases. Accordingly, in a specific range of temperature (550–750 °C), calcination affects the output of synthesized zeolites and its crystallinity, but not the zeolite types.

3.3.4. Crystallization Temperature

Crystallization is the process that builds the crystal framework of zeolites and that occurs faster at higher temperatures. According to the related literature [31], the crystalline phase will transfer to other types when the temperature is over a specific value. Therefore, a series of experiments about crystallization temperatures were carried out, and the products were analyzed by XRD, as given in Figure 10. It is shown that there are no distinct peaks of crystal phases at room temperature (25 °C), which indicates that low temperature is not suitable for zeolites formation. As temperature rises to about 90 °C, the ZX crystals gradually show up and grow rapidly, reaching a plateau at 120 °C. Nevertheless, zeolite HS occurs and replaces zeolite X bit by bit when the temperature is beyond 120 °C. Quite mounts of zeolite HS were formed when the crystallization temperature is up to 150 °C. This behavior is not occasional. Mostly, faujasite type zeolites have a larger specific surface area and pore size than hydroxysodalite. So, it is an unstable phase in the reaction system, tending to convert into other stable types when the condition changes [31]. In addition, the yield increases as the temperature increases. It is inferred that 120 °C is a relatively suitable crystallization temperature for ZX formation.

3.4. Synthesis Mechanism

Acid pretreatment removes the undesirable Fe2O3 located in the surface of the basalt particle, enhancing the relative content of Si2O3 and Al2O3 indirectly. In the alkali fusion stage, OH spreads into the system, surrounding the basalt particles and permeating into the channels. The reactive phase was decomposed gradually, resulting in the concentration of Al3+ and Si4+ increasing. However, the dissociation course would not last all the time and reach an equilibrium as the alkalinity descends and as the Na-aluminate and Na-silicate salts covered the surface of basalt particle. During this fusion period, no zeolite crystals are occurring, as shown in Figure 11, which is consistent with the liquid phase transition mechanism.
This process can be described as follows:
N a O H + x A l 2 O 3 y S i O 2 F u s i o n N a 2 S i O 3 + N a 2 A l O 2
In the aging stage, Na+, Al3+ and Si4+ were dissolved into aqueous solution. Stirring facilitates the release of Al3+ and Si4+. As we know that Si4+ reacts readily with Al3+ precursors to generate aluminosilicates compounds. When enough quantity of silicate ions was generated in the reaction system, aluminates and silicates are condensed to form an aluminosilicate gel onto the surface of basalt particles, as depicted in the equation:
N a O H ( a q ) + N a 2 A l ( O H ) 4 ( a q ) + N a 2 S i O 3 ( a q ) A g i n g [ N a x ( A l O 2 ) ( S i O 2 ) 2 N a O H H 2 O ] ( g e l )
In the crystallization stage, crystallization of zeolites takes place through nucleation reaction and crystal growth. In alkaline conditions, aluminum constitutes negatively charged tetrahedral species, a structure coinciding with the zeolite framework. A previous work shows that there exists an incubation period of 3h during which the nucleation occurs [31]. Thereafter, zeolite X crystal grows rapidly for a few hours. However, ZX is an unstable phase in comparison to hydroxysodalite. Once beyond some critical point, the equilibrium will be interrupted, resulting in a collapse in the ZX framework and being rearranged into HS crystal. The period may be demonstrated with the reaction equation, as below:
[ N a x ( A l O 2 ) ( S i O 2 ) 2 N a O H H 2 O ] ( g e l ) C r y s t a l l i z a t i o n N a p [ ( A l O 2 ) p ( S i O 2 ) q h H 2 O ]

4. Conclusions

(a) Compared to TH, AFH is apracticalapproach for zeolite preparation from basalt. Zeolite X can be synthesized with high purity and crystallinity by using AFH.
(b) Acid treatment, alkali/basalt ratio, calcination temperature and crystallization temperature play a significant role in the conversion from basalt to zeolites. The acid treatment enhances the yield and quality of synthetic zeolites; calcination temperature (550–750°C) affects the degree of crystallinity, but not the synthesized zeolite types. NaOH/basalt ratio and crystallization temperature affect not only the degree of zeolite crystallinity but also the synthesized zeolite types.
(c) The quality and type of the synthesized zeolites could vary significantly, depending on the formation conditions and parameters. The optimum synthesis condition of zeolite X was acid treatment of 5wt% HCl solution, NaOH/basalt ratio of 1:1, calcination temperature of 650 °C and crystallization temperature of 120 °C.

Author Contributions

P.Y. and G.K. conceived and designed the study; H.S. performed the experiment and analyzed the data; P.Y. provided critical feedback and helped shape the research; H.S., P.Y. wrote the final version of the manuscript.

Funding

This research was funded bythe Natural Science Foundation of Hunan Province (Grant No. 2019JJ60003).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jae, J.; Tompsett, G.A.; Foster, A.J.; Hammond, K.; Auerbach, S.M.; Lobo, R.F.; Huber, G.W. Investigation into the shape selectivity of zeolite catalysts for biomass conversion. J. Catal. 2011, 279, 257–268. [Google Scholar] [CrossRef]
  2. Moliner, M.; Román-Leshkov, Y.; Davis, M.E. Tin-containing zeolites are highly active catalysts for the isomerization of glucose in water. Proc. Natl. Acad. Sci. USA 2010, 107, 6164–6168. [Google Scholar] [CrossRef] [PubMed]
  3. Primo, A.; Garcia, H. Zeolites as catalysts in oil refining. Chem. Soc. Rev. 2014, 43, 7548–7561. [Google Scholar] [CrossRef] [PubMed]
  4. Al-Nasser, A.Y.; Al-Zenki, S.F.; Al-Saffar, A.E.; Abdullah, F.K.; Mashaly, M.; Al-Bahouh, M. Zeolite as a feed additive to reduce Salmonella and improve production performance in broilers. Int. J. Poult. Sci. 2011, 10, 448–454. [Google Scholar] [CrossRef]
  5. Eroglu, N.; Emekci, M.; Athanassiou, C.G. Applications of natural zeolites on agriculture and food production. J. Sci. Food Agric. 2017, 97, 3487–3499. [Google Scholar] [CrossRef]
  6. Wang, S.; Peng, Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chem. Eng. J. 2010, 156, 11–24. [Google Scholar] [CrossRef]
  7. Xu, Y.; Ke, G.; Yin, J.; Lei, W.; Yang, P. Synthesis of thiol-functionalized hydrotalcite and its application for adsorption of uranium (VI). J. Radioanal. Nucl. Chem. 2019, 319, 791–803. [Google Scholar] [CrossRef]
  8. Liu, Y.; Yang, P.; Li, Q.; Liu, Y.; Yin, J. Preparation of FeS@Fe3O4 core-shell magnetic nanoparticles and their application in uranyl ions removal from aqueous solution. J. Radioanal. Nucl. Chem. 2019, 321, 499–510. [Google Scholar] [CrossRef]
  9. Barth-Wirsching, U.; Holler, H. Experimental studies on zeolite formation conditions. Eur. J. Miner. 1989, 489–506. [Google Scholar] [CrossRef]
  10. Gougazeh, M.; Buhl, J.C. Synthesis and characterization of zeolite A by hydrothermal transformation of natural Jordanian kaolin. J. Assoc. Arab Univ. Basic Appl. Sci. 2014, 15, 35–42. [Google Scholar] [CrossRef]
  11. Qian, T.; Li, J. Synthesis of Na-A zeolite from coal gangue with the in-situ crystallization technique. Adv. Powder Technol. 2015, 26, 98–104. [Google Scholar] [CrossRef]
  12. Ojumu, T.V.; Du Plessis, P.W.; Petrik, L.F. Synthesis of zeolite A from coal fly ash using ultrasonic treatment–A replacement for fusion step. Ultrason. Sonochemistry 2016, 31, 342–349. [Google Scholar] [CrossRef] [PubMed]
  13. Shaban, M.; Abukhadra, M.R.; Shahien, M.G.; Ibrahim, S.S. Novel bentonite/zeolite-NaP composite efficiently removes methylene blue and Congo red dyes. Environ. Chem. Lett. 2018, 16, 275–280. [Google Scholar] [CrossRef]
  14. Shawabkeh, R.; Al-Harahsheh, A.; Hami, M.; Khlaifat, A. Conversion of oil shale ash into zeolite for cadmium and lead removal from wastewater. Fuel 2004, 83, 981–985. [Google Scholar] [CrossRef]
  15. Wajima, T.; Haga, M.; Kuzawa, K.; Ishimoto, H.; Tamada, O.; Ito, K.; Nishiyama, T.; Downs, R.T.; Rakovan, J.F. Zeolite synthesis from paper sludge ash at low temperature (90 °C) with addition of diatomite. J. Hazard. Mater. 2006, 132, 244–252. [Google Scholar] [CrossRef]
  16. Moisés, M.P.; da Silva, C.T.P.; Meneguin, J.G.; Girotto, E.M.; Radovanovic, E. Synthesis of zeolite NaA from sugarcane bagasse ash. J. Mater. Lett. 2013, 108, 243–246. [Google Scholar]
  17. Zhao, Y.; Zhang, B.; Zhang, X.; Wang, J.; Liu, J.; Chen, R. Preparation of highly ordered cubic NaA zeolite from halloysite mineral for adsorption of ammonium ions. J. Hazard. Mater. 2010, 178, 658–664. [Google Scholar] [CrossRef]
  18. Wajima, T.; Ikegami, Y. Synthesis of zeolitic materials from waste porcelain at low temperature via a two-step alkali conversion. J. Ceramics international, 2007, 33(7): 1269-1274. J. Ceram. Int. 2007, 33, 1269–1274. [Google Scholar] [CrossRef]
  19. Lin, G.; Zhuang, Q.; Cui, Q.; Wang, H.; Yao, H. Synthesis and adsorption property of zeolite FAU/LTA from lithium slag with utilization of mother liquid. J. Chin. J. Chem. Eng. 2015, 23, 1768–1773. [Google Scholar] [CrossRef]
  20. Derouane, E.G.; Determmerie, S.; Gabelica, Z.; Blom, Z. Synthesis and characterization of ZSM-5 type zeolites I. physico-chemical properties of precursors and intermediates. Appl. Catal. 1981, 1, 201–224. [Google Scholar] [CrossRef]
  21. Behin, J.; Bukhari, S.S.; Kazemian, H.; Rohani, S. Developing a zero liquid discharge process for zeolitization of coal fly ash to synthetic NaP zeolite. Fuel 2016, 171, 195–202. [Google Scholar] [CrossRef]
  22. Hollman, G.G.; Steenbruggen, G.; Janssen-Jurkovičová, M. A two-step process for the synthesis of zeolites from coal fly ash. Fuel 1999, 78, 1225–1230. [Google Scholar] [CrossRef]
  23. Pal, P.; Das, J.K.; Das, N.; Bandyopadhyay, S. Synthesis of NaP zeolite at room temperature and short crystallization time by sonochemical method. Ultrason. Sonochemistry 2013, 20, 314–321. [Google Scholar] [CrossRef] [PubMed]
  24. Wajima, T.; Ikegami, Y. Synthesis of crystalline zeolite-13X from waste porcelain using alkali fusion. Ceram. Int. 2009, 35, 2983–2986. [Google Scholar] [CrossRef]
  25. Marakatti, V.S.; Rao, P.V.C.; Choudary, N.V.; Ganesh, G.S.; Shah, G.; Maradur, S.P.; Halgeri, A.B.; Shanbhag, G.V.; Ravishankar, R. Influence of Alkaline Earth Cation Exchanged X-Zeolites Towards Ortho-Selectivity in Alkylation of Aromatics: Hard-Soft-Acid-Base Concept. Adv. Porous Mater. 2014, 2, 221–229. [Google Scholar] [CrossRef]
  26. Ma, Y.; Yan, C.; Alshameri, A.; Qiu, X.; Zhou, C.; Li, D. Synthesis and characterization of 13X zeolite from low-grade natural kaolin. Adv. Powder Technol. 2014, 25, 495–499. [Google Scholar] [CrossRef]
  27. Chen, D.; Hu, X.; Shi, L.; Cui, Q.; Wang, H.; Yao, H. Synthesis and characterization of zeolite X from lithium slag. Appl. Clay Sci. 2012, 59, 148–151. [Google Scholar] [CrossRef]
  28. Purnomo, C.W.; Salim, C.; Hinode, H. Synthesis of pure Na–X and Na–A zeolite from bagasse fly ash. J. Microporous Mesoporous Mater. 2012, 162, 6–13. [Google Scholar] [CrossRef]
  29. Cardoso, A.M.; Paprocki, A.; Ferret, L.S.; Azevedo, C.M.N.; Pires, M. Synthesis of zeolite Na-P1 under mild conditions using Brazilian coal fly ash and its application in wastewater treatment. Fuel 2015, 139, 59–67. [Google Scholar] [CrossRef]
  30. Murayama, N.; Yamamoto, H.; Shibata, J. Mechanism of zeolite synthesis from coal fly ash by alkali hydrothermal reaction. Int. J. Miner. Process. 2002, 64, 1–17. [Google Scholar] [CrossRef]
  31. Doyle, A.M.; Alismaeel, Z.T.; Albayati, T.M.; Abbas, A.S. High purity FAU-type zeolite catalysts from shale rock for biodiesel production. Fuel 2017, 199, 394–402. [Google Scholar] [CrossRef]
Figure 1. Characterization patterns of raw basalt powder (sample A). (a) X-ray diffractogram; (b) N2 adsorption-desorption isotherm; (c) and (d) SEM images.
Figure 1. Characterization patterns of raw basalt powder (sample A). (a) X-ray diffractogram; (b) N2 adsorption-desorption isotherm; (c) and (d) SEM images.
Materials 12 03895 g001
Figure 2. X-ray diffractograms of sample A and synthesized products of sample A, B and D by using TH under the condition: NaOH concentration of 5 M, liquid/solid ratio of 5:1, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Figure 2. X-ray diffractograms of sample A and synthesized products of sample A, B and D by using TH under the condition: NaOH concentration of 5 M, liquid/solid ratio of 5:1, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Materials 12 03895 g002
Figure 3. X-ray diffractogram of synthetic product of sample C by AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Figure 3. X-ray diffractogram of synthetic product of sample C by AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Materials 12 03895 g003
Figure 4. SEM images of synthesized product of sample C by using AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Figure 4. SEM images of synthesized product of sample C by using AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Materials 12 03895 g004
Figure 5. FTIR spectrum of zeolite X synthesized by AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Figure 5. FTIR spectrum of zeolite X synthesized by AFH under the condition: NaOH/basalt ratio of 1:1, calcination temperature of 650 °C, aging time of 10 h, crystallization temperature of 120 °C, crystallization time of 12 h.
Materials 12 03895 g005
Figure 6. N2 adsorption-desorption isotherm of synthesized zeolite X.
Figure 6. N2 adsorption-desorption isotherm of synthesized zeolite X.
Materials 12 03895 g006
Figure 7. X-ray diffractograms of synthetic products from sample A and sample C by AFH under the same condition.
Figure 7. X-ray diffractograms of synthetic products from sample A and sample C by AFH under the same condition.
Materials 12 03895 g007
Figure 8. Diffractograms of products synthesized by using AFH with different NaOH/basalt ratios (Designation: L: labradorite; A: anorthite; ZX: zeolite X; HS: hydroxylsodalite).
Figure 8. Diffractograms of products synthesized by using AFH with different NaOH/basalt ratios (Designation: L: labradorite; A: anorthite; ZX: zeolite X; HS: hydroxylsodalite).
Materials 12 03895 g008
Figure 9. Diffractograms of products synthesized by using AFH under different calcination temperatures.
Figure 9. Diffractograms of products synthesized by using AFH under different calcination temperatures.
Materials 12 03895 g009
Figure 10. Diffractograms of products synthesized by using AFH under different crystallization temperatures.
Figure 10. Diffractograms of products synthesized by using AFH under different crystallization temperatures.
Materials 12 03895 g010
Figure 11. X-ray diffractogram of sample E.
Figure 11. X-ray diffractogram of sample E.
Materials 12 03895 g011
Table 1. Chemical composition of raw basalt powder.
Table 1. Chemical composition of raw basalt powder.
ComponentsSiO2Al2O3Fe2O3CaOMgONa2OK2OSO3Ignition LossTotal
wt.%47.8918.1714.675.614.32.101.500.063.1897.48
Table 2. Surface area and pore-structure parameters of raw basalt powder.
Table 2. Surface area and pore-structure parameters of raw basalt powder.
SampleSurface Area (m2·g−1)Mircopore Volume (t-Plot) (cm3·g−1)Pore Volume (cm3·g−1)Pore Size (nm)
Sample A12.11960.0027500.0192976.36874
Table 3. Surface area and pore-structure parameters of synthesized zeolite X.
Table 3. Surface area and pore-structure parameters of synthesized zeolite X.
SampleSurface Area (BET) (m2·g−1)Mircopore Volume (t-Plot) (cm3·g−1)Pore Volume (cm3·g−1)Pore Size (nm)
Zeolite X486.46020.1719090.2708442.22706
Table 4. Main chemical composition of basalt powder after acid treatment (sample C).
Table 4. Main chemical composition of basalt powder after acid treatment (sample C).
ComponentsSiO2Al2O3Fe2O3
wt%59.6524.358.23

Share and Cite

MDPI and ACS Style

Ke, G.; Shen, H.; Yang, P. Synthesis of X-Zeolite from Waste Basalt Powder and its Influencing Factors and Synthesis Mechanism. Materials 2019, 12, 3895. https://doi.org/10.3390/ma12233895

AMA Style

Ke G, Shen H, Yang P. Synthesis of X-Zeolite from Waste Basalt Powder and its Influencing Factors and Synthesis Mechanism. Materials. 2019; 12(23):3895. https://doi.org/10.3390/ma12233895

Chicago/Turabian Style

Ke, Guojun, Haichen Shen, and Pengfei Yang. 2019. "Synthesis of X-Zeolite from Waste Basalt Powder and its Influencing Factors and Synthesis Mechanism" Materials 12, no. 23: 3895. https://doi.org/10.3390/ma12233895

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop