Lithium extraction and hydroxysodalite zeolite synthesis by

lithium grease lubricants, flux additive for aluminum production, and lithium ion batteries. (LIBs).1 ...... stabilisation of arsenic from arsenic-ric...
0 downloads 0 Views 761KB Size
Subscriber access provided by UNIV OF WATERLOO

Article

Lithium extraction and hydroxysodalite zeolite synthesis by hydrothermal conversion of #-spodumene Peng Xing, Chengyan Wang, Lei Zeng, Baozhong Ma, Ling Wang, Yongqiang Chen, and Cheng Yang ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.9b00923 • Publication Date (Web): 06 May 2019 Downloaded from http://pubs.acs.org on May 7, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Lithium extraction and hydroxysodalite zeolite synthesis by hydrothermal conversion of αspodumene Peng Xing, Chengyan Wang,* Lei Zeng, Baozhong Ma,* Ling Wang, Yongqiang Chen, and Cheng Yang School of Metallurgical and Ecological Engineering, University of Science and Technology Beijing, Beijing 100083, China. Mailing address: 30 Xueyuan Road, Haidian District, Beijing 100083, China Corresponding Authors: [email protected] (C. Wang) and [email protected] (B. Ma).

ABSTRACT: The continuously increasing demand for lithium has made it one of the strategic metals, rendering its exploitation of critical importance. Natural α-spodumene is still the primary resource of lithium extraction. The traditional process for the treatment of α-spodumene generates immense quantities of waste residue and needs a high-temperature heat treatment, leading to high energy consumption. In addition to lithium, α-spodumene is rich in aluminum and silicon, and thus it is a potential raw material for zeolite synthesis. Herein, a novel process was developed for the clean and efficient extraction of lithium from α-spodumene coupled with the synthesis of hydroxysodalite zeolite. By hydrothermal alkaline treatment, α-spodumene was

ACS Paragon Plus Environment

1

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 25

converted into hydroxysodalite; the lithium in α-spodumene was released into the solution, and subsequently recovered by precipitation with Na2CO3. A lithium extraction efficiency of 95.8% was obtained under the optimum conditions: temperature 250 °C, NaOH concentration 600 g/L, liquid/solid ratio 5:1, stirring speed 500 rpm, and reaction time 2 h. In addition, the influences of various factors on the composition and textural properties of the product were analyzed using XRD, SEM, TG, N2 adsorption/desorption, and FTIR.

KEYWORDS: α-Spodumene, Lithium, Extraction, Hydroxysodalite, Synthesis INTRODUCTION Lithium and its compounds have been widely applied in heat-resistant glass and ceramics, lithium grease lubricants, flux additive for aluminum production, and lithium ion batteries (LIBs).1 As the most promising rechargeable battery, LIB has become the largest consumer of lithium.2 The unprecedented growth in electric vehicles has dramatically increased the demand for lithium. A shortage of lithium carbonate, one of the most important raw materials for LIB production, is projected to occur around 2020.3 The high economic importance of lithium has made it one of the strategic metals. In recent years, much attention and research has been devoted to the extraction of lithium from brine.

4-8

In comparison, there are few literature reports on the extraction of lithium from α-

spodumene. Natural α-spodumene is a lithium aluminum inosilicate mineral, which is the important mineral source for lithium extraction and generally has a high content of Li2O (~5%). Since the middle of the twentieth century, α-spodumene has been processed by a traditional method which involves a high-temperature heat treatment (~1100 °C) to transform the inert αspodumene into active β-spodumene, followed by sulfuric acid curing at around 250 °C and

ACS Paragon Plus Environment

2

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

water leaching.9-12 Although some direct leaching processes have been reported,13-16 heat treatment is still a necessary step. The phase transformation of α-spodumene involves high energy consumption. In addition, the processing of 1 t of α-spodumene normally generates approximately 0.8 t of leaching residue using the traditional process. The treatment of immense quantities of this lithium leaching residue has become an increasingly serious problem. These have severely affected the sustainable exploitation of α-spodumene. Zeolites are microporous, aluminosilicate minerals commonly used in desiccation, catalysis, and waste water treatment.17-19 Industrially important zeolites are synthetically produced by a hydrothermal reaction of alumina, silicon, and sodium hydroxide. The searching for low-cost and easily available raw materials for the synthesis of zeolites has been a constant research focus.20,21 Natural raw materials have economic advantages over synthetic chemicals.22 Hydroxysodalite (Na8[AlSiO4]6(OH)2∙nH2O) is a mineral of the zeolite group, made of a cubic array of β-cages and exhibits a structure similar to that of sodalite.22,23 Hydroxysodalite is hydrophilic and has numerous micro and meso channels, which enables it to be used for the separation of small molecules from gas or liquid mixtures, hydrogen storage, and catalysis;24,25 it has also been used to modify mortar.26 So far, much work has focused on the synthesis of hydroxysodalite using low-cost resources containing alumina and silicon, such as kaolinite and coal fly ash.27-32 In addition, Kato et al. synthesized hydroxysodalite using the alkaline waste solution of glass polishing powder.33 Esaifan et al. synthesized hydroxysodalite using the basalt powder.24 In fact, in addition to lithium, α-spodumene is rich in aluminum and silicon, and thus it is a potential raw material for zeolite synthesis. However, there is no report on the conversion of α-spodumene to zeolite by hydrothermal treatment.

ACS Paragon Plus Environment

3

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 25

In this work, a novel process was proposed for the clean and efficient extraction of lithium from α-spodumene coupled with the synthesis of hydroxysodalite zeolite. By hydrothermal alkaline treatment α-spodumene was converted to hydroxysodalite while the lithium in αspodumene was released into the solution and then recovered as Li2CO3 by adding Na2CO3. The process has the advantages of lower energy consumption and no gas emission. Additionally, the synthesized hydroxysodalite zeolite is a value-added product. The influences of various factors on the lithium extraction and the composition and textural properties of the product were investigated in detail. EXPERIMENTAL SECTION Materials. The α-spodumene ore was crushed and ground to a particle size below 0.074 mm. The chemical composition of the α-spodumene ore is shown in Table 1. Sodium hydroxide, cetyltrimethyl ammonium bromide (CTAB), calcium oxide, and sodium carbonate were of analytically pure grade. Table 1. Chemical Composition of the α-Spodumene Ore

content (wt %)

Li

Al

K

Fe

Ca

Na

Mg

SiO2

2.57

14.3

2.73

1.45

1.09

0.36

0.23

57.0

Apparatuses and Procedure. The hydrothermal alkaline treatment of α-spodumene was performed in a vertical autoclave. The α-spodumene ore (80 g) was first mixed with NaOH solution at a certain liquid-solid ratio in a liner. The slurry was agitated with electric-motor driven impellers. The temperature was controlled by regulating the heating and the cooling water. After the hydrothermal alkaline treatment was over, the slurry was filtered to separate the lithium solution and solid product. The solid product was washed with deionized water and oven-dried. The desilication was also carried out in the autoclave and the stirring rate was fixed

ACS Paragon Plus Environment

4

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

at 500 rpm. The precipitation of lithium was performed in beaker with mechanical stirring at 500 rpm in a water bath. Analysis Methods. Lithium in α-spodumene ore and alkaline treatment products was analyzed by inductively coupled plasma optical emission spectrometry (ICP-OES) (Optima 7000DV, PerkinElmer). The silicon in ore and solution was analyzed by silicon molybdenum blue spectrophotometry. X-ray diffraction (XRD) measurements were carried out using a Rigaku Ultima IV diffractometer (Cu, Kα). The surface morphology and composition of the αspodumene and the products of alkaline treatment were analyzed using field-emission scanning electron microscope (FESEM; JSM-7001F) and energy dispersive X-ray spectroscope (EDS; BRUKER XFlash5010). Thermogravimetric (TG) measurement was performed using a NETZSCH STA 409C analyzer under a nitrogen atmosphere at a heating rate of 10 °C/min. The functional groups of the products of alkaline treatment were analyzed using an iN10MX FTIR spectrometer. The textural characterization of the products was obtained from low-temperature N2 adsorption/desorption at -196 °C on an ASAP 2020 Micromeritics instrument with an equilibrium time of 10 min. The specific surface area was estimated by the Brunauer-EmmetTeller (BET) method. The micropore volume was determined by the t-plot method. Total pore volume was estimated by the single point adsorption at P/P0 = 0.98. The pore size distribution was calculated by the Berrett-Joyner-Halenda (BJH) method. RESULTS AND DISCUSSION Characterization of the α-Spodumene Ore. The XRD pattern (Figure 1) and SEM−EDS analysis (Figure S1, Supporting Information) of the α-spodumene ore indicate that α-spodumene (LiAlSi2O6) is the major phase, with quartz (SiO2) and muscovite (KAl2Si3AlO10(OH)2) as minor

ACS Paragon Plus Environment

5

ACS Sustainable Chemistry & Engineering

phases. The cuboid shape α-spodumene shows a smooth surface with a size of ~20 μm in width and ~30 μm in length (Figure S1). 

16000 

Intensity (Counts)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

-Spodumene (PDF#33-0786)

Quartz (PDF#46-1045) Muscovite 2M1 (PDF#07-0032)





12000

 



8000  

4000











   



  



 





 

 





0 20

40

60

80

2 

Figure 1. XRD pattern of the α-spodumene ore. Hydrothermal Conversion of α-Spodumene and Lithium Extraction. Effect of Temperature on Lithium Extraction and Product Composition. Hydrothermal treatment is an effective method for mineral dissociation and conversion.34-36 Temperature is an important factor in hydrothermal treatment. Figure 2a shows the lithium extraction vs. temperature curve under the conditions of 600 g/L NaOH concentration, 5:1 liquid/solid ratio, 2 h reaction time, and 500 rpm stirring rate. It was observed that the lithium extraction increased with the increase of temperature. The extraction ratio of lithium reached a plateau at 250 °C. Thus the optimum temperature was selected to be 250 °C. As shown in Figure 3, the diffraction peak of α-spodumene remained unchanged after hydrothermal treatment at 150 °C, which explains the low extraction efficiency of lithium. A new phase (hydroxysodalite, Na4Al3Si3O12(OH)) peak appeared at 200 °C. The peak of αspodumene decreased whereas the diffraction peak of hydroxysodalite markedly increased with the increase of temperature. When the temperature was increased to 250 °C, hydroxysodalite

ACS Paragon Plus Environment

6

Page 7 of 25

became the main phase. Meanwhile, the maximum extraction yield of lithium was obtained. A small amount of an impurity, zeolite (1.2Na2O·0.8CaO·Al2O3·2SiO2·H2O) was derived from the reaction of the calcium, aluminum, and silicon in the ore with the alkaline solution. Figure 3 indicates that the conversion of α-spodumene in the alkaline solution can occur at above 200 °C and that increasing temperature can accelerate the conversion. The following reaction is predicted to take place during the hydrothermal conversion of α-spodumene: 6LiAlSi2O6 + 14NaOH → 2Na4Al3Si3O12(OH) + 3Li2SiO3 + 3Na2SiO3 + 6H2O 100

(a)

100

80

60

40

(b)

60

40

20

20

0 140

(1)

80

Li extraction (%)

Li extraction (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

0

160

180

200

220

240

260

280

300

400

500

600

NaOH concentration (g/L)

Temperature (C)

ACS Paragon Plus Environment

7

ACS Sustainable Chemistry & Engineering

(c)

100

100

(d)

80

Li extraction (%)

Li extraction (%)

80

60

40

60

40

20

20

0

0 3:1

4:1

5:1

200

6:1

400

600

800

Stirring rate (rpm)

Liquid/solid ratio

(e)

100

Li extraction (%)

80

60

40

20

0 0

1

2

3

4

5

Time (h)

Figure 2. Effects of (a) temperature, (b) NaOH concentration, (c) liquid/solid ratio, (d) stirring rate, and (e) leaching time on the extraction of lithium.  -Spodumene  Quartz  Muscovite 2M1 

 Hydroxysodalite (PDF#11-0401)

250 C

 Zeolite (PDF#18-1210)

















  

   

 

 

  

 

 



Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 25

 











230 C  



 









 



200 C

 

 



























  







  







 

20



 



150 C

 



 

 



raw ore 40











60







80

2 

ACS Paragon Plus Environment

8

Page 9 of 25

Figure 3. XRD patterns of the α-spodumene ore converted at various temperatures. Effect of NaOH Concentration on Lithium Extraction and Product Composition. According to eq 1, NaOH concentration is also a key factor for extraction. Therefore, the variations in lithium extraction with NaOH concentration were investigated. The temperature, liquid/solid ratio, reaction time, and stirring rate were maintained at 250 °C, 5:1, 2 h, and 500 rpm, respectively. As shown in Figure 2b, the effect of the NaOH concentration on the extraction of lithium was significant. When the NaOH concentration was 600 g/L, the extraction of lithium reached a plateau and hence the optimum NaOH concentration was selected to be 600 g/L. As shown in Figure 4, the peak of α-spodumene decreased, while a new phase (Li2SiO3) peak appeared and the faujasite (Na14Al12Si13O51·6H2O) became the dominant phase in product after the hydrothermal treatment at NaOH concentration of 300 g/L. Interestingly, when the NaOH concentration was increased to 400 g/L, hydroxysodalite replaced faujasite as the dominant phase. Accordingly, the transformation of faujasite into hydroxysodalite strongly depends on the NaOH concentration. This result is consistent with the transformation of zeolite A into hydroxysodalite.37 The intensities of the α-spodumene and Li2SiO3 peaks decreased with increasing NaOH concentration, which accounts for the increase in lithium extraction.  -Spodumene  Li2SiO3 (PDF#29-0828)





Faujasite-Na (PDF#28-1036)  Zeolite (PDF#18-1210)

 Hydroxysodalite (PDF#11-0401)





 









NaOH 600 g/L







   









 



 





Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering











  







20

 



   





    





 



 



NaOH 500 g/L











     



  



NaOH 400 g/L







   



NaOH 300 g/L 

40

 

 



60

2 

ACS Paragon Plus Environment

9

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

Figure 4. XRD patterns of the α-spodumene ore converted at various NaOH concentrations. Effect of Liquid/Solid Ratio on Lithium Extraction. The experiments were performed at different liquid/solid ratios (3:1–6:1) while the temperature, NaOH concentration, stirring rate, and reaction time were set at 250 °C, 600 g/L, 500 rpm, and 2 h, respectively. Figure 2c shows that the extraction efficiency of lithium initially increased with the increase of liquid/solid ratio. Increasing liquid/solid ratio promoted the mass transfer between the ore and the NaOH solution and increased the amount of NaOH, which was favorable to the conversion of α-spodumene. Thus, based on Figure 2c, the optimum liquid/solid ratio was selected to be 5:1. Figures 2a−2c indicate that the conditions for the conversion of α-spodumene into hydroxysodalite are more stringent than those for kaolinite.27,38,39 Effect of Stirring Rate on Lithium Extraction. Increasing stirring rate can also promote mass transfer. Hence, the effect of agitation on lithium extraction was investigated at different stirring speeds that ranged from 200 to 700 rpm. The temperature, NaOH concentration, liquid/solid ratio, and reaction time were maintained at 250 °C, 600 g/L, 5:1, and 2 h, respectively. As shown in Figure 2d, 500 rpm was sufficient for Li extraction. Effect of Reaction Time on Lithium Extraction and Composition and Textural Properties of Product. Numerous literatures have shown that the reaction time is an important factor affecting the synthesis of zeolites. In the case of the synthesis of hydroxysodalite, it usually takes several or many hours.23,29,30,40 Thus, the influence of reaction time on the Li extraction and the composition and textural properties of the product was investigated for durations from 0.5 to 5 h, under the constant conditions of 250 °C, 600 g/L NaOH concentration, 5:1 liquid/solid ratio, and 500 rpm stirring rate. It was observed that the hydrothermal conversion of α-spodumene was fast, with a Li extraction efficiency of above 70% within 0.5 h (Figure 2e), and a large amount of

ACS Paragon Plus Environment

10

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

hydroxysodalite was found in the hydrothermal treatment product (Figures 5 and 6). The peaks of α-spodumene decreased with time and finally disappeared after 2 h (Figure 5). This observation explains the results in Figure 2e. As shown in Figure 6a, the hydroxysodalite shows a regular cubic structure and the particle size of which is about 20 μm. The surface of hydroxysodalite became rough and its particle size increased with increasing reaction time (Figure 6a, b, and d). Figure 7 shows the N2 adsorption/desorption isotherms for the products obtained under different hydrothermal conditions. According to the IUPAC classification, all the products display the typical reversible curve of Type II isotherm, which represents unrestricted monolayer-multilayer adsorption.41 This finding is consistent with the results obtained by Esaifan et al.24 Pore size distributions are shown in Figure 8, which indicates that the synthesized products present a porous structure. With the reaction time increasing from 0.5 to 5 h, the average pore diameter decreased from 5.92 to 4.82 nm. The textural properties of various products are listed in Table S1 (Supporting Information). The BET surface area and pore volume of the synthesized products increased with the reaction duration, which accounts for the much higher N2 adsorption capacity of the products obtained at 2 and 5 h (Figure 7). The Li extraction reached equilibrium within 2 h. The further time extension did not significantly improve the extraction or change the composition of the product. However, the intensity of the diffraction peak of hydroxysodalite did increase with time, indicating the increase in crystallinity of the product. Previous studies on the synthesis of hydroxysodalite from kaolin by hydrothermal alkaline treatment have shown that prior to the crystallization of hydroxysodalite, intermediate metastable phases, faujasite and zeolite A were generally formed at the initial stage of the reaction.32,42 Nevertheless, for the hydrothermal conversion of α-

ACS Paragon Plus Environment

11

ACS Sustainable Chemistry & Engineering

spodumene to hydroxysodalite, the formation of faujasite was mainly related to the NaOH concentration rather than to the reaction time, which implies that hydroxysodalite was directly formed from the precursor and the aluminosilicate gel.29 Figure 9 shows the FTIR spectra of the products synthesized at different reaction times. The peaks at 3543 and 3418 cm−1 are attributed to the –OH vibration of structural water molecules in hydroxysodalite. The peaks at 1650 and 1533 cm−1 correspond to the OH deformation of water. The pronounced wide absorption band at 885 cm−1 corresponds to the asymmetric stretching vibration of T–O–T (T = Si, Al). The absorption band at 727 and 661 cm−1 correspond to the symmetric stretching of T–O–T. The absorption band at 488 cm−1 is the characteristic of the bending vibration of O–T–O. The characteristics of the FTIR spectra of the products are consistent with those of hydroxysodalite.22,40,44 The weakening of the –OH and T–O–T vibrations and the disappearance of the O–T–O vibration were observed when the reaction time was increased from 2 to 5 h.  -Spodumene



 Hydroxysodalite (PDF#11-0401)

 Zeolite (PDF#18-1210)

















  



5h 

  

 

 

  

 

 

 

 

 

  





Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 25



 

  



 

2h











  



1h



   

  

 

  

20

 



0.5 h



     



40

60

80

2 

Figure 5. XRD patterns of the α-spodumene ore converted at various times.

ACS Paragon Plus Environment

12

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 6. SEM images and EDX mapping of the hydroxysodalite produced from α-spodumene at different reaction time and dosage of CTAB: (a) 0.5 h; (b) 2 h; (c) 2 h with 3 g/L CTAB; (d) 5 h; (e) 5 h with 3 g/L CTAB (constant conditions: 250 °C, 600 g/L NaOH concentration, 5:1 liquid/solid ratio, and 500 rpm stirring rate).

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering

0.5 h 2h 2 h with CTAB 5h 5 h with CTAB

15

3

Quantity adsorbed (cm /g STP)

20

10

5

0 0.0

0.2

0.4

0.6

0.8

1.0

Relative pressure (P/P0)

Figure 7. N2 adsorption/desorption isotherms of the products synthesized under different hydrothermal conditions. 0.005 0.5 h 2h 2 h with CTAB 5h 5 h wiht CTAB

0.004

3

dV/dw Pore Volume (cm /(gnm))

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

0.003

0.002

0.001

0.000 2.5

5.0

7.5

10.0

12.5

15.0

Pore Width (nm)

Figure 8. Pore size distribution for the products synthesized under different hydrothermal conditions. An integrated experiment was performed under the established optimized conditions (250 °C, 600 g/L NaOH concentration, 5:1 liquid/solid ratio, 2 h reaction time, and 500 rpm stirring rate) and a 95.8% Li extraction yield was obtained. The thermo-gravimetric analysis data of the hydroxysodalite zeolite synthesized at the optimized conditions revealed a 5.5% mass loss from room temperature to 1000 °C (Figure S2,

ACS Paragon Plus Environment

14

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Supporting Information). The mass loss from room temperature to 150 °C was due to the removal of adsorbed water while the larger mass loss from 400 to 700 °C was due to the dehydroxylation.22,43 Effect of the CTAB Surfactant on Lithium Extraction and Composition and Textural Properties of Product. Some literatures have been reported the promotion of cationic surfactants on minerals dissociation and crystal formation.45,46 Hence, a common cationic surfactant, CTAB was used in the present work to study its effects on lithium extraction and the composition and properties of the product. The dosage of CTAB was 3 g/L. Other parameters of hydrothermal reaction included a temperature of 250 °C, a NaOH concentration of 600 g/L, a liquid/solid ratio of 5:1, a stirring rate of 500 rpm, and a time of 2 h. It was observed that CTAB had no obvious effect on the lithium extraction. Interestingly, a small amount of a new phase, potassium aluminum oxide silicate (K0.85Al0.85Si0.15O2), was found in the product synthesized in the presence of CTAB (Figure S3, Supporting Information). The BET surface area of product increased with the addition of CTAB in the hydrothermal synthesis. This finding is in agreement with previous research results.47-49 However, CTAB had a small effect on the pore volume, which may be related to its relatively low concentration. Additionally, adding CTAB was conductive to obtaining near-spherical particles of a uniform size (Figure 6e). As shown in Figure 9, the addition of CTAB had little effect on the FTIR spectra of the products; only the weakening of the T–O–T vibration at 727 cm−1 was observed.

ACS Paragon Plus Environment

15

727

Page 16 of 25

488

661

885

5 h with CTAB

1650 1533

3543

Transmission (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

3418

ACS Sustainable Chemistry & Engineering

5h

2h

4000

3500

3000

2500

2000

1500

1000

500

-1

Wavenumber (cm )

Figure 9. FTIR spectra of the products synthesized under different hydrothermal conditions. Desilication. The silicon in α-spodumene was partially leached in the hydrothermal treatment. The dissolved silica crystallizes easily, thereby causing the decrease in purity of lithium carbonate. To avoid it, the liquor containing lithium was further purified via desilication using inexpensively available calcium oxide to remove the silica. The desilication product was calcium silicate, which has many applications.50 In order to facilitate desilication, an equal volume of deionized water was added to dilute the liquor. After dilution, the concentration of SiO2 in solution was 36.4 g/L. Based on our previous research on desilication,51 the mass ratio of CaO to SiO2 and the time for desilication were set at 1.2 and 1 h, respectively. Only the effect of the reaction temperature on desilication was investigated in the present work. Figure S4 (Supporting Information) shows the SiO2 precipitation vs. desilication temperature curve. The extent of desilication was enhanced with increasing temperature, because the elevated temperature was conducive to the nucleation and growth of calcium silicate. As shown in Figure S4, a temperature of 95 °C was sufficient for desilication and a SiO2 precipitation of 96.2% was obtained.

ACS Paragon Plus Environment

16

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Lithium Precipitation. In order to enrich the Li in solution to facilitate precipitation, the solution after desilication was concentrated by evaporation, replenished with NaOH and then returned to the hydrothermal alkaline treatment step to leach raw ore. A continuous loop of desilication, evaporation concentration, and hydrothermal alkaline treatment was carried out and a concentrated lithium solution (Li 10.6 g/L) was obtained. Based on the widely used method of preparing lithium carbonate in industry,52 a stoichiometric amount of saturated Na2CO3 solution was added to 200 mL solution. The temperature and time for precipitation were 95 °C and 1 h, respectively. A lithium precipitation of 67.1% was achieved under these conditions. Although the direct precipitation rate of lithium is relatively low it would not affect the overall recovery rate since the solution is recycled. The XRD pattern of the obtained lithium carbonate (Figure S5, Supporting Information) agrees well with the standard pattern. CONCLUSIONS Lithium extraction and hydroxysodalite zeolite synthesis were successfully achieved by the hydrothermal conversion of α-spodumene. The optimum reaction temperature, NaOH concentration, liquid/solid ratio, reaction time, and stirring rate for Li extraction were determined to be 250 °C, 600 g/L, 5:1, 2 h, and 500 rpm, respectively. A 95.8% Li extraction yield was obtained under these conditions. After desilication, the lithium in solution was recovered by precipitation with Na2CO3. The conversion of α-spodumene occurred in the alkaline solution at 200 °C and was basically completed at 250 °C. At the low NaOH concentration of 300 g/L, the main product of the hydrothermal conversion was faujasite, which was transformed into hydroxysodalite with the further increase of the NaOH concentration. The hydrothermal conversion of α-spodumene was fast, as a large concentration of hydroxysodalite was present after only 0.5 h. The synthesized

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

products presented a porous structure. Their BET surface area, pore volume, and average pore diameter increased with the duration of the treatment. Additionally, the FTIR spectra of the synthesized products indicates that the weakening of the –OH and T–O–T vibrations and the disappearance of the O–T–O vibration took place when the reaction time was increased from 2 to 5 h. The addition of CTAB had no obvious effect on the lithium extraction but increased the BET surface area of the products. Furthermore, adding CTAB was conductive to obtaining nearspherical particles with a uniform size. ASSOCIATED CONTENT Supporting Information Additional results as shown in Figures S1−S5, and Table S1 (PDF) AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected]. Tel.: +86-10-62332271. *E-mail: [email protected]. ORCID Chengyan Wang: 0000-0003-3982-6208 Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS

ACS Paragon Plus Environment

18

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

This research was funded by the National Natural Science Foundation of China (U1802253 and 51674026), Beijing Natural Science Foundation (2182040), and the National Key R&D Program of China (2018YFC1900304). REFERENCES (1) Jaskula, B. W. Mineral Commodity Summaries Lithium; U.S. Geological Survey, Reston, VA, 2018; pp 98–99. (2) Swain, B. Recovery and recycling of lithium: A review. Sep. Purifi. Technol. 2017, 172, 388–403. (3) Choubey, P. K.; Kim, M. S.; Srivastava, R. R.; Lee, J. C.; Lee, J. Y. Advance review on the exploitation of the prominent energy-storage element: Lithium. Part I: From mineral and brine resources. Miner. Eng. 2016, 89, 119–137. (4) Flexer, V; Baspineiro, C. F.; Galli, C. I. Lithium recovery from brines: A vital raw material for green energies with a potential environmental impact in its mining and processing. Sci. Total Environ. 2018, 639, 1188–1204. (5) Ooi, K.; Sonoda, A.; Makita, Y.; Chitrakar, R.; Tasaki-Handa, Y.; Nakazato, T. Recovery of lithium from salt-brine eluates by direct crystallization as lithium sulfate. Hydrometallurgy 2017, 174, 123–130. (6) Paranthaman, M. P.; Li, L.; Luo, J.; Hoke, T.; Ucar, H.; Moyer, B. A.; Harrison, S. Recovery of lithium from geothermal brine with lithium–aluminum layered double hydroxide chloride sorbents. Environ. Sci. Technol. 2017, 51, 13481–13486.

ACS Paragon Plus Environment

19

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

(7) Xu, X.; Chen, Y.; Wan, P.; Gasem, K.; Wang, K.; He, T.; Adidharma, H.; Fan, M. Extraction of lithium with functionalized lithium ion-sieves. Prog. Mater. Sci. 2016, 84, 276–313. (8) Wang, Y.; Liu, H.; Fan, J.; Liu, X.; Hu, Y.; Hu, Y.; Zhou, Z.; Ren, Z. Recovery of lithium ions from salt lake brine with a high Mg/Li ratio using heteropoly acid ionic liquid. ACS Sustainable Chem. Eng. 2019, 7, 3062–3072. (9) Peltosaari, O.; Tanskanen, P.; Heikkinen, E. P.; Fabritius, T. α→γ→β-phase transformation of spodumene with hybrid microwave and conventional furnaces. Miner. Eng. 2015, 82, 54–60. (10) Salakjani, N. K.; Singh, P.; Nikoloski, A. N. Mineralogical transformations of spodumene concentrate from Greenbushes, Western Australia. Part 1: Conventional heating. Miner. Eng. 2016, 98, 71–79. (11) Lajoie-Leroux, F.; Dessemond, C.; Soucy, G.; Laroche, N.; Magnan, J. F. Impact of the impurities on lithium extraction from β-spodumene in the sulfuric acid process. Miner. Eng. 2018, 129, 1–8. (12) Ellestad, R. B.; Leute, K. M. Method of extracting lithium values from spodumene ores, 1950, US patent 2,516,109. (13) Kuang, G.; Liu, Y.; Li, H.; Xing, S.; Li, F.; Guo, H. Extraction of lithium from βspodumene using sodium sulfate solution. Hydrometallurgy 2018, 177, 49–56. (14) Rosales, G. D.; Ruiz, M. D. C.; Rodriguez, M. H. Novel process for the extraction of lithium from β-spodumene by leaching with HF. Hydrometallurgy 2014, 147–148, 1–6. (15) Chen, Y.; Tian, Q.; Chen, B.; Shi, X.; Liao, T. Preparation of lithium carbonate from spodumene by a sodium carbonate autoclave process. Hydrometallurgy 2011, 109, 43–46.

ACS Paragon Plus Environment

20

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(16) Rosales, G. D.; Ruiz, M. C.; Rodriguez, M. H. Study of the extraction kinetics of lithium by leaching β-Spodumene with hydrofluoric acid. Minerals 2016, 6, 98. (17) Bai, L.; Li, K.; Yan, Y.; Jia, X.; Lee, J.; Yang, Y. Catalytic epoxidation of cis-cyclooctene over vanadium-exchanged faujasite zeolite catalyst with ionic liquid as cosolvent. ACS Sustainable Chem. Eng. 2016, 4, 437–444. (18) Wang, S.; Peng, Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chem. Eng. J. 2010, 156, 11–24. (19) Bailey, S. E.; Olin, T. J.; Bricka, R. M.; Adrian, D. D. A review of potentially low-cost sorbents for heavy metals. Wat. Res. 1999, 33, 2469–2479. (20) Nascimento, M.; Soares, P. S. M.; Souza, V. P. D. Adsorption of heavy metal cations using coal fly ash modified by hydrothermal method. Fuel 2009, 88, 1714–1719. (21) Li, X.; Jiang, Y.; Liu, X.; Shi, L.; Zhang, D.; Sun, L. Direct synthesis of zeolites from a natural clay, attapulgite. ACS Sustainable Chem. Eng. 2017, 5, 6124–6130. (22) Gaidoumi, A. E.; Benabdallah, A. C.; Bali, B. E.; Kherbeche, A. Synthesis and characterization of zeolite HS using natural pyrophyllite as new clay source. Arab. J. Sci. Eng. 2018, 43, 191–197. (23) Kundu, D.; Dey, B.; Naskar, M. K.; Chatterjee, M. Emulsion-derived urchin-shaped hydroxy sodalite particles. Mater. Lett. 2010, 64, 1630–1633. (24) Esaifan, M.; Hourani, M.; Khoury, H.; Rahier, H.; Wastiels, J. Synthesis of hydroxysodalite zeolite by alkali-activation of basalt powder rich in calc-plagioclase. Adv. Powder Technol. 2017, 28, 473–480.

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

(25) Kazemimoghadam, M.; Mohammadi, T. Preparation of nano pore hydroxysodalite zeolite membranes using of kaolin clay and chemical sources. Desalination 2011, 278, 438–442. (26) Sasnauskas, V.; Palubinskaitė, D. The synthesis of hydrosodalite and its use in mortar technology. Mater. Sci-Poland 2005, 23, 793–801. (27) Temuujin, J.; Okada, K.; MacKenzie, K. J. D. Zeolite formation by hydrothermal treatment of waste solution from selectively leached kaolinite. Mater. Lett. 2002, 52, 91–95. (28) Marsh, A.; Heath, A.; Patureau, P.; Evernden, M.; Walker, P. A mild conditions synthesis route to produce hydrosodalite from kaolinite, compatible with extrusion processing. Micropor. Mesopor. Mat. 2018, 264, 125–132. (29) Fukui, K.; Kanayama, K.; Yamamoto, T.; Yoshida, H. Effects of microwave irradiation on the crystalline phase of zeolite synthesized from fly ash by hydrothermal treatment. Adv. Powder Technol. 2007, 18, 381–393. (30) Woolard, C. D.; Strong, J.; Erasmus, C. R. Evaluation of the use of modified coal ash as a potential sorbent for organic waste streams. Appl. Geochem. 2002, 17, 1159–1164. (31) Golbad, S.; Khoshnoud, P.; Abu-Zahra, N. Hydrothermal synthesis of hydroxyl sodalite from fly ash for the removal of lead ions from water. Int. J. Environ. Sci. Technol. 2017, 14, 135–142. (32) Gualtieri, A.; Norby, P.; Artioli, G.; Hanson, J. Kinetic study of hydroxysodalite formation from natural kaolinites by time-resolved synchrotron powder diffraction. Micropor. Mat. 1997, 9, 189–201.

ACS Paragon Plus Environment

22

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(33) Kato, K.; Yoshioka, T.; Okuwaki, A. Study for recycling of ceria-based glass polishing powder II-Recovery of hydroxysodalite from the alkali waste solution containing SiO2 and Al2O3. Ind. Eng. Chem. Res. 2000, 39, 4148–4151. (34) Rubisov, D. H.; Papangelakis, V. G. Sulphuric acid pressure leaching of laterites-a comprehensive model of a continuous autoclave. Hydrometallurgy 2000, 58, 89–101. (35) Xue, N.; Zhang, Y.; Liu, T.; Huang, J.; Zheng, Q. Effects of hydration and hardening of calcium sulfate on muscovite dissolution during pressure acid leaching of black shale. J. Clean. Prod. 2017, 149, 989–998. (36) Chen, Y.; Liu, N.; Ye, L.; Xiong, S.; Yang, S. A cleaning process for the removal and stabilisation of arsenic from arsenic-rich lead anode slime. J. Clean. Prod. 2018, 176, 26–35. (37) Subotić, B.; Škrtic, D.; Šmit, I.; Sekovanić, L. Transformation of zeolite A into hydroxysodalite I. An approach to the mechanism of transformation and its experimental evaluation. J. Cryst. Growth 1980, 50, 498–508. (38) Alkan, M.; Hopa, Ç.; Yilmaz, Z.; Güler, H. The effect of alkali concentration and solid/liquid ratio on the hydrothermal synthesis of zeolite NaA from natural kaolinite. Micropor. Mesopor. Mat. 2005, 86, 176–184. (39) Querol, X.; Moreno, N.; Umaña, J. C.; Alastuey, A.; Hernández, E.; López-Soler, A.; Plana, F. Synthesis of zeolites from coal fly ash: an overview. Int. J. Coal Geol. 2002, 50, 413– 423. (40) Naskar, M. K.; Kundu, D.; Chatterjee, M. Effect of process parameters on surfactantbased synthesis of hydroxy sodalite particles. Mater. Lett. 2011, 65, 436–438.

ACS Paragon Plus Environment

23

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

(41) Thommes, M.; Kaneko, K.; Neimark, A. V.; Olivier, J. P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K. S. W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. (42) Covarrubias, C.; García, R.; Arriagada, R.; Yánez, J.; Garland, M. T. Cr(III) exchange on zeolites obtained from kaolin and natural mordenite. Micropor. Mesopor. Mat. 2006, 88, 220– 231. (43) Khajavi, S.; Sartipi, S.; Gascon, J.; Jansen, J. C.; Kapteijn, F. Thermostability of hydroxy sodalite in view of membrane applications. Micropor. Mesopor. Mat. 2010, 132, 510–517. (44) Majdinasab, A. R.; Yuan, Q. Microwave synthesis of zeolites from waste glass cullet using indirect fusion and direct hydrothermal methods: A comparative study. Ceram. Int. 2019, 45, 2400–2410. (45) Barik, R.; Sanjay, K.; Mishra, B. K.; Mohapatra, M. Micellar mediated selective leaching of manganese nodule in high temperature sulfuric acid medium. Hydrometallurgy 2016, 165, 44– 50. (46) Smith, D. K.; Korgel, B. A. The importance of the CTAB surfactant on the colloidal seedmediated synthesis of gold nanorods. Langmuir 2008, 24, 644–649. (47) Wang, X.; Chen, H.; Meng, F.; Gao, F.; Sun, C.; Sun, L.; Wang, S.; Wang, L.; Wang, Y. CTAB resulted direct synthesis and properties of hierarchical ZSM-11/5 composite zeolite in the absence of template. Micropor. Mesopor. Mat. 2017, 243, 271–280.

ACS Paragon Plus Environment

24

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(48) Singhala, A.; Gangwar, B. P.; Gayathry, J. M. CTAB modified large surface area nanoporous geopolymer with high adsorption capacity for copper ion removal. Appl. Clay Sci. 2017, 150, 106–114. (49) Sun, L.; Wang, Y.; Chen, H.; Sun, C.; Meng, F.; Gao, F.; Wang, X. Direct synthesis of hierarchical ZnZSM-5 with addition of CTAB in a seeding method and improved catalytic performance in methanol to aromatics reaction. Catal. Today 2018, 316, 91–98. (50) Ismail, H.; Shamsudin, R.; Hamid, M. M. A. Effect of autoclaving and sintering on the formation of β-wollastonite. Mat. Sci. Eng. C 2016, 58, 1077–1081. (51) Xing, P.; Wang, C.; Wang, L.; Ma, B.; Chen, Y.; Wang, G. Clean and efficient process for the extraction of rubidium from granitic rubidium ore. J. Clean. Prod. 2018, 196, 64–73. (52) Zhang, J.; Hu, J.; Liu, Y.; Jing, Q.; Yang, C.; Chen, Y.; Wang, C. Sustainable and facile method for the selective recovery of lithium from cathode scrap of spent LiFePO4 batteries. ACS Sustainable Chem. Eng. 2019, 7, 5626−5631. Table of Contents graphic

Synopsis The lithium extraction and hydroxysodalite zeolite synthesis have been successfully achieved by hydrothermal conversion of α-spodumene.

ACS Paragon Plus Environment

25