Structures, formation mechanisms, and ion-exchange properties of α-, β-, and γ-Na2TiO3

Fancheng Mengabc, Yahui Liu*ab, Tianyan Xueab, Qian Subc, Weijing Wangab and Tao Qi*ab
aNational Engineering Laboratory for Hydrometallurgical Cleaner Production Technology, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China. E-mail: yhliu@ipe.ac.cn; tqgreen@home.ipe.ac.cn
bKey Laboratory of Green Process and Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
cUniversity of Chinese Academy of Sciences, Beijing 100039, China

Received 1st July 2016 , Accepted 17th November 2016

First published on 21st November 2016


Abstract

α-, β-, and γ-Na2TiO3 were prepared from rutile TiO2 and molten NaOH. Three models of β-Na2TiO3 with space groups of R[3 with combining macron], P[1 with combining macron], and P[3 with combining macron] were proposed, and the R[3 with combining macron] model was refined from the experimental data by using the Rietveld method. The structure of β-Na2TiO3 is a superstructure of α-Na2TiO3 and supposedly contains Ti6O19 clusters. The structures of Na2TiO3 were mainly determined by the particle sizes of rutile and the reaction temperatures. α-Na2TiO3 could be prepared from fine rutile particles (D(50) < 25.8 μm) and molten NaOH at 500 °C or quenching the melt of Na2TiO3 at 1000 °C quickly. γ- and β-Na2TiO3 were the thermodynamically stable phases of Na2TiO3 at around 500 °C and above 800 °C, respectively. α-Na2TiO3 was formed far beyond the thermodynamically stable state. The Na+ in α-Na2TiO3 was easier to exchange with H+ in water than that in β or γ phases. They all converted to amorphous phases after the 2nd, 6th, and 4th water washings at 25 °C, respectively. β-Na2TiO3 followed similar paths of ion-exchange as α-Na2TiO3, which was different from that of γ-Na2TiO3.


1. Introduction

Various kinds of sodium titanates have been prepared for use as cation exchangers,1,2 photocatalysts,3,4 and precursors to prepare TiO2 nanotubes.5,6 Most researches have focused on the nanostructures of titanates consisting of negatively charged layered sheets constructed with TiO6 octahedra and obtained by a hydrothermal method.7,8 Moreover, sodium titanates can also be prepared from the reaction of Na2CO3 and TiO2.9,10 There are fewer studies of Na2TiO3 (sodium meta-titanate) compared to other sodium titanates. However, Na2TiO3 is receiving more attention because of its possible applications for the adsorption of radioactive nuclides,11,12 CO2 capture at high temperatures,13 electrolysis for producing titanium,14 and preparation of H2TiO3–lithium adsorbents.15 Furthermore, Na2TiO3 hydrolyzes upon contact with water, which could generate NaOH.16–18 This makes Na2TiO3 an ideal intermediate for producing TiO2 or upgrading titania slag by a process that involves alkaline roasting of titanium minerals and water washing for NaOH recycling.19–24 As an example, our group used this property of Na2TiO3 to develop the molten NaOH process for TiO2 production.24

According to the phase diagrams of the Na2O–TiO2 system shown in Fig. 1,25 Na2TiO3 could be obtained from the system of Na2O–TiO2 with a TiO2 content of 40–60 wt% at 400–1000 °C. α-, β-, and γ-Na2TiO3 were prepared from the reaction of TiO2 and Na2CO3 or NaOH;26–28 however, these preparation methods are not very reproducible. α-Na2TiO3 exhibits the NaCl-type structure (space group Fm[3 with combining macron]m, Z = 4/3); Na and Ti atoms occupy the same sites in the unit cell with 2/3 and 1/3 probability, respectively.17,29 γ-Na2TiO3 has a chain structure consisting of the rarely observed trigonal-bipyramidal TiO5 units (space group C2/c, Z = 4).28 The XRD pattern of β-Na2TiO3 reported as PDF 00-047-0130 (a = 3.195 Å, c = 7.676 Å) was attributed to the rhombohedrally distorted NaCl-type subcell;30 however, the superlattice reflections were wrongly omitted from this card in the PDF database. The corrected pattern appeared as PDF 00-050-110 (space group R[3 with combining macron]; a = b = 13.927 Å, c = 7.676 Å; Z = 19). β-Na2TiO3 was primarily treated as a monoclinic form, indexed as PDF 00-037-0345,26 however, the rhombohedral indexing was more reliable than the monoclinic one.31 However, 19 is not a multiple of 3, as is necessary for the rhombohedral lattice. Therefore, some Na/Ti disorder was suggested,31 and it was refined in this study.


image file: c6ra16984h-f1.tif
Fig. 1 Phase diagrams for the Na2O–TiO2 system: (a) TiO2 content <60%, and (b) TiO2 content >60%.25

There are some misunderstandings related to the structures of Na2TiO3. For instance, Na6Ti2O7 in Fig. 1 should be γ-Na2TiO3.26 Other kinds of sodium titanates prepared by a hydrothermal method were wrongly identified as Na2TiO3 in some studies.32,33 A sodium titanate coating formed on the titanium surface by NaOH treatment was thought to be Na2TiO3.34 However, the X-ray diffraction (XRD) patterns could not be indexed to the three phases discussed above; a likely structure is that of Na2Ti5O11.35 Additionally, the space group of α-Na2TiO3 was mistaken for Immm.13

The properties among the three kinds of Na2TiO3 structures were found to be different. In the molten NaOH process for TiO2 production as mentioned above, the crystal structures of Na2TiO3 affected the Ti yield coefficient, and α-Na2TiO3 was the ideal phase that could be prepared from fine natural rutile.21 However, the reaction mechanisms of molten NaOH and TiO2 in forming α-, β-, and γ-Na2TiO3 remained unclear, especially for the transformations among the three phases.26,27 Moreover, many further applications of Na2TiO3 must consider the ion-exchange properties, e.g., the molten NaOH process for TiO2 production need Na2TiO3 with good ion-exchange ability for the recycle of NaOH medium. Thus, a comparison of the ion-exchange properties in water should be studied.

In this study, α-, β-, and γ-Na2TiO3 were successfully prepared from the reaction of molten NaOH and rutile TiO2. The structure of β-Na2TiO3 was refined. The effects of particle sizes of rutile and reaction temperatures were studied to reveal the formation and transformation mechanisms of α-, β-, and γ-Na2TiO3. Their ion-exchange properties in water were investigated for comparison.

2. Experimental

2.1 Materials

Analytical-grade NaOH, rutile TiO2, and deionized water were used throughout the experiment. Natural rutile concentrate (claret colour, Hubei province, China) was used as a substitute for rutile TiO2 at different particle sizes. The chemical composition (weight fraction) of the natural rutile was 90.62% TiO2, 4.01% SiO2, 1.49% ∑Fe, 0.54% MgO, and 1.16% Al2O3. The main phase of natural rutile was rutile TiO2.

2.2 Experimental details

NaOH pellets were ground into powders in a mortar made of WC–Co (tungsten carbide 84.5 wt% and cobalt 15.5 wt%) with the protection of N2 in the glove-box. NaOH powder and rutile TiO2 were mixed in the mortar with a molar ratio of 2.2[thin space (1/6-em)]:[thin space (1/6-em)]1. The mixture was put into the nickel crucibles that were then heated in a muffle furnace. α- and γ-Na2TiO3 were prepared by heating the mixture of NaOH and TiO2 at 500 °C for 90 min. β-Na2TiO3 was prepared at 800 °C for 90 min and cooled slowly to room temperature. Excessive NaOH was used to guarantee the complete reaction. In preparing α- and β-Na2TiO3, analytical-grade TiO2 powder with an average particle size (D(50)) of 0.5 μm was used. γ-Na2TiO3 was prepared from the coarse TiO2 particles with secondary particle sizes of about 100 μm, obtained from prilling the above TiO2 powder.

After the reaction, the products were ground and washed with ethanol to remove a part of excess NaOH in order to avoid the moisture of the samples. γ-Na2TiO3 sample was contaminated by a small amount of α-Na2TiO3 which was converted to an amorphous form by additional washes with ethanol.

In the ion-exchange tests, α-, β-, and γ-Na2TiO3 were washed with water in a 250 mL flask heated by an oil bath with a liquid/solid mass ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]1 at 25 °C and 60 °C. Each washing lasted for 30 min, and the pulp was filtered after each washing. Then, the filter cake was sampled and dried for further analysis. The removal of Na+ in the water washing was used to characterize the ion-exchange abilities of α-, β-, and γ-Na2TiO3 in water.

2.3 Characterisation

XRD patterns were obtained using an X-ray diffractometer (X'Pert PRO MPD, PANalytical, Netherlands), recorded within 5–90° 2θ using CuKα radiation. The phase semi-quantification of Na2TiO3 was obtained using the GSAS-EXPGUI software packages with the Rietveld method.36,37 In phase semi-quantification of Na2TiO3, the zero-point correction, unit cell parameters, scale factor, background, and peak width were refined, while the atomic coordinates, occupancies, and thermal parameters were fixed. The weight fraction (wi) for each phase was obtained from the refinement relation:
 
image file: c6ra16984h-t1.tif(1)
where i is a particular phase among the N phases, Si is the refined scale factor, Z is the number of formula units per cell, M is the formula mass, and V is the volume of the unit cell.38

The particle size distributions of the TiO2 precursors were measured using a laser particle analyser (Mastersizer 2000, Malvern Instruments, UK).

The reactions of NaOH and rutile TiO2 were investigated in a microfurnace by Raman spectrometer (LABRAM HR800, Horiba Jobin Yvon) in situ. It was equipped with an intensive charge-coupled device (ICCD), and the pulsed exciting light (514.5 nm) from a Q-switch pulsed THG-Nd:YAG laser was focused using an Olympus BH-2 microscope. To investigate the ion-exchange process of Na2TiO3, the resulting slurry after the water washing was placed on a small glass plate to collect the Raman spectra.

The particle morphologies and elemental compositions of the products were characterized using a scanning electron microscope (SEM; JSM-7001F, Electron Company, Japan) coupled to energy dispersive X-ray spectroscopy (EDS; INCA X-MAX, Oxford Instruments, US).

Thermogravimetry (TG) and differential scanning calorimetry (DSC) experiments (STA 449C, NETZSCH, Germany) were performed from 25 °C to 900 °C in a Pt crucible.

The chemical compositions of the samples and solutions were examined by inductively coupled plasma atomic emission spectroscopy (ICP-OES; Optima 5300DV, PerkinElmer, USA).

3. Results and discussion

3.1 Crystal structures and characterisation of Na2TiO3

As mentioned in the introduction, the structures of α- and γ-Na2TiO3 have been well studied. α- and γ-Na2TiO3 were successfully prepared in this study. The XRD patterns after lattice refinement and ball-stick models (created by the VESTA39) of α- and γ-Na2TiO3 are shown in Fig. 2. All XRD peaks were well indexed.
image file: c6ra16984h-f2.tif
Fig. 2 XRD patterns after the refinement and ball-stick models (inset) of α-Na2TiO3 (a), β-Na2TiO3 (b, space group of R[3 with combining macron]), and γ-Na2TiO3 (c) (red, O; yellow, Na; blue, Ti).

As suggested by V. B. Nalbandyan,31 there may be Na/Ti mixing in some positions of β-Na2TiO3 with the space group of R[3 with combining macron]. The original atomic coordinates were obtained from the procedures shown in Fig. S1 by Density Functional Theory (DFT) calculations and revising the atomic site of Na (0, 0, 0) by 2/3 Na and 1/3 Ti. Lattice parameter refinement was done firstly, and the obtained simple R[3 with combining macron] model (its atomic sites were shown in Table S1) was used to do structural refinement and create the P[1 with combining macron] and P[3 with combining macron] models in the following text. Then structural refinement on the XRD data in Fig. 2b was done. The atomic coordinates, occupancies, and thermal parameters were all optimized. The results were shown in Tables 1 and 2. The structural residual factors from Rietveld refinement were small. In Fig. 2b, all XRD peaks were well indexed with this R[3 with combining macron] model, and its Miller indices were added.

Table 1 Crystallographic data of β-Na2TiO3 after refinement
Formula Na38Ti19O57
Space group R[3 with combining macron]
Crystal system Trigonal
Lattice parameters a = b = 13.9252(5) Å, c = 7.7024(3) Å
Gaussian (G) and Lorentzian (L) profile coefficients GU = 0, GV = 45.48, GW = 18.76, LX = LY = 1.00
Reliability factors Rp = 7.03%, Rwp = 9.48%, χ2 = 2.426


Table 2 Atomic positions, occupancies, and thermal parameters of β-Na2TiO3 with space group of R[3 with combining macron] after refinement
Atom Site x/a y/b z/c Occ. Uiso
O1 18f 0.6796 0.4540 0.5257 1.0000 0.0525
O2 18f 0.0799 0.8491 0.5019 1.0000 0.0173(7)
O3 18f 0.2234 0.9642 0.7946 1.0000 0.0120(4)
O4 3b 0.6667 0.3333 0.8333 1.0000 0.0252
Na1 18f 0.4253 0.3669 0.0110 0.9444 0.0251(1)
Ti1 18f 0.4253 0.3669 0.0110 0.0556 0.0251(7)
Na2 18f 0.1442 0.1238 0.3326 0.1111 0.0276(7)
Ti2 18f 0.1442 0.1238 0.3326 0.8889 0.0276(8)
Na3 3a 0.0000 0.0000 0.0000 0.3333 0.0306
Ti3 3a 0.0000 0.0000 0.0000 0.6667 0.0305(5)
Na4 18f 0.1089 0.8467 0.0109 1.0000 0.0106(6)


As shown in Table 2, there are another two Na/Ti mixing positions in the unit cell except for (0, 0, 0). The ball-stick model of β-Na2TiO3 (space group R[3 with combining macron]) in Fig. 2b indicates that it is not a framework or layered structure, and it contains Ti6O19 clusters some of which are connected into Ti13O38 dimers via edge sharing with the Ti2 octahedron. However, the Ti2 site is partially substituted by Na2 and this precludes formation of infinite oxotitanate chains along the c direction. When the clusters are stacked according to the face centred lattice, the NaCl-type crystal structure will be obtained. Thus, β-Na2TiO3 can be considered as a superstructure of α-Na2TiO3 which is in accord with the prediction of V. B. Nalbandyan.31

Table 3 shows the average bond lengths of Na/Ti and O atoms in β-Na2TiO3. The ionic radii sums of six-coordination Na(+1) and O(−2), Ti(+4) and O(−2) are 2.42 and 2.005 Å, respectively.40 The average bonds lengths in Table 3 were all close to the predicted values. According to the Bond Valence model,41,42 the central Na distributes its valence for six neighboured O atoms, and the calculated Na–O bond length is 2.47 Å, which is very close to the bond length of Na4–O (2.40 Å). Similarly, the calculated Ti–O bond length for the central six-coordination O atom from Bond Valence model is 2.22 Å. For the O4 atom, it is surrounded by six Na2/Ti2 mixing positions with the Ti occupancy of 0.889. The bond length of Na2/Ti2–O4 is 2.28 Å, a little larger to the calculated value. The structure is highly disordered due to substitution of cations that differ significantly in size. Then, positions of nominally identical anions contacting with them should also be disordered, and some tabulated distances should not be taken for real interatomic distances.

Table 3 The average bond lengths of Na/Ti and O atoms in β-Na2TiO3 after refinement (Å)
Atom 1 Atom 2 Average bond length Prediction from ionic radii sumsa
a Prediction = 2.42 × occupancy(Na) + 2.005 × occupancy(Ti).
Na1/Ti1 O 2.39 2.40
Na2/Ti2 O 2.02 2.05
Na3/Ti3 O 2.18 2.14
Na4 O 2.40 2.42


To obtain the β-Na2TiO3 models with definite atomic sites (the need for DFT calculations to study its properties in future studies), another two models with space groups P[1 with combining macron] and P[3 with combining macron] were obtained by geometric transformation from the rhombohedral and hexagonal cell of the R[3 with combining macron] model with the 2/3 Na and 1/3 Ti at site (0, 0, 0) in Table S1. For this simple R[3 with combining macron] model of β-Na2TiO3, the bond valence calculations in Tables S2 and S3 indicated that this structure was reasonable. For the P[1 with combining macron] model, the lattice parameters are Z = 19, a = b = 8.440 Å, c = 19.689 Å, α = 85.23°, β = 94.77°, and γ = 68.83°. For the P[3 with combining macron] model, Z = 57, a = b = 13.925 Å, c = 23.107 Å. Their atomic sites obtained from DFT calculations are shown in Tables S4 and S5. Thus, these two models were not directly proved by the Rietveld analysis which should be done in the future study.

The morphologies of α-, β-, and γ-Na2TiO3 were slightly different. α-Na2TiO3 had a more porous surface and a larger specific surface area (see Fig. S2 and S3).

Fig. 3 shows the Raman spectra of α-, β-, and γ-Na2TiO3, and they are in agreement with a previous report.27 α- and β-Na2TiO3 have similarities in their Raman spectra. However, the poor disordered structure because of the random distribution of Na and Ti atoms endows α-Na2TiO3 with very broad Raman peaks.27 β- and γ-Na2TiO3 have clearly different and numerous sharply resolved peaks. The peaks at 500–1000 cm−1 are mostly connected to Ti–O–Ti stretching vibrations, and the peak at 800–815 cm−1 can be ascribed to the difference in Ti–O bond lengths.43 The band at approximately 1080 cm−1 is related to the CO32− symmetric stretching vibration from Na2CO3 that remained in the three samples.44 The peaks at approximately 454 and 657 cm−1 are assigned to the Ti–O–Ti stretching in the edge-shared TiO6 units in β-Na2TiO3, while the peaks at approximately 176 and 296 cm−1 are attributed to the Na–O–Ti bending modes.45,46 For γ-Na2TiO3, the peaks at approximately 385 and 632 cm−1 indicate the edge-shared TiO5 units.47


image file: c6ra16984h-f3.tif
Fig. 3 Raman spectra of the prepared α-, β-, and γ-Na2TiO3.

3.2 Formation and transformation of α-, β-, and γ-Na2TiO3

3.2.1. Effect of the particle sizes of rutile TiO2. To investigate the effect of the particle sizes of the rutile TiO2 precursor, natural rutile which were milled to D(50) = 2.9, 9.1, 25.9, 58.5, 67.5, 76.2, 85.3, and 103.0 μm substituted for rutile TiO2. The XRD patterns of the products and the quantification analysis of the α-Na2TiO3 phase in each product are shown in Fig. 4. When the D(50) of natural rutile was smaller than 25.9 μm, the α-Na2TiO3 phase formed, and when D(50) > 58.5 μm, γ-Na2TiO3 was the main phase of the product with about 15 wt% of α-Na2TiO3.
image file: c6ra16984h-f4.tif
Fig. 4 XRD patterns (a) and weight fraction of α-Na2TiO3 (b) in the products obtained from natural rutile with different average particle sizes.

Raman spectra were collected continuously during the reactions of NaOH and rutile of different average particle sizes, as shown in Fig. 5. The rutile crystals belong to the space group of P42/mnm with Raman active modes occurring around 143 cm−1 (B1g), 445 cm−1 (Eg), 612 cm−1 (A1g), and 243 cm−1 (second order Raman scattering) in Fig. 5.48 As shown in Fig. 5a and b, the Eg mode shifted to lower frequencies, the band at 243 cm−1 shifted to higher frequencies, while the B1g and A1g modes remained unchanged with the increasing temperature (<380 °C). The Raman measurements were performed at ambient conditions; thus, NaOH could readily absorb CO2 from the air and form Na2CO3. The intense band at 1081–1074 cm−1 is attributed to Na2CO3,44 and the bands shifted to lower frequencies with increasing temperatures. In both Fig. 5a and b, the Raman active modes of rutile slowly disappeared with the appearance of the bands of α- and γ-Na2TiO3 at 806 and 798 cm−1, respectively. Blue shifts occurred at above 380 °C compared to the Raman spectra at room temperature (Fig. 3). During the reaction of NaOH and rutile with different average particle sizes, no intermediate phases were observed during the formation of the α- and γ-Na2TiO3 phases.


image file: c6ra16984h-f5.tif
Fig. 5 In situ Raman spectra of the mixture of NaOH and rutile from 30 °C to 500 °C. Natural rutile (a) D(50) = 2.9 μm and (b) D(50) = 76.2 μm.

TG/DSC measurements were used to study the details of the reactions with rutile of different particle sizes. The TG graphs in Fig. 6 show a continuous weight loss at about 300–600 °C. This weight loss can be related to the release of water from the reaction. The end of this weight loss indicated the end of the reaction. Fig. 6 indicates that fine rutile requires less time to complete the reaction, which means that fine rutile particles could improve the reaction rate. There are only two obvious endothermic peaks at about 62 °C and 282 °C in the DSC curve, corresponding to the evaporation of free water and the melt of the NaOH + Na2CO3 eutectic mixture,49 respectively. No other exothermic or exothermal peaks were observed below 600 °C. Thus, it is reasonable to conclude that there is no crystal transformation between α- and γ-Na2TiO3 in the reactions below 600 °C.


image file: c6ra16984h-f6.tif
Fig. 6 TG/DSC curves of the mixture of NaOH and natural rutile from 25 °C to 900 °C. Natural rutile (a) D(50) = 2.9 μm and (b) D(50) = 76.2 μm.

The XRD patterns of the product obtained at different reaction time are shown in Fig. 7. The main peak of rutile at approximately 27.5° is close to the main peak of γ-Na2TiO3 at 27.6–27.8°. This shows that the rutile phase decreased with the increased reaction time, and γ-Na2TiO3 formed gradually. A very small amount of α-Na2TiO3 formed at the beginning of the reaction, and its amount remained nearly constant in the reaction. Thus, the crystal transformation did not occur between α- and γ-Na2TiO3 in the reaction, which was in agreement with Fig. 6.


image file: c6ra16984h-f7.tif
Fig. 7 XRD patterns of the products obtained from natural rutile D(50) = 76.2 μm for different reaction time.

The product was formed by inter-diffusion upon thermal activation.50 Fig. 8a is backscattered SEM of the intermediate product after 10 min being embedded in the resin and polished, while Fig. 8b is the general SEM of the same intermediate product. The surface of the rutile particle reacted with NaOH at the beginning of the reaction, while the interior part of the particle remained unchanged (see EDS analysis in Fig. 8c and d). A compact product layer consisting of the sodium titanate formed and wrapped the surface of the rutile particle.


image file: c6ra16984h-f8.tif
Fig. 8 Backscattered SEM (a) and general SEM (b) of the intermediate product at 10 min prepared from natural rutile D(50) = 85.3 μm; (c) and (d) EDS analysis of spot ① and ② in (a) and (b), respectively.

In conclusion, at 500 °C, the fine rutile particles of D(50) < 25.8 μm reacted with NaOH very fast, thus favoured the formation of α-Na2TiO3 in which Na and Ti atoms were distributed randomly. Conversely, the product layer formed on the surface of rutile of larger particle sizes made the diffusion of NaOH slow and difficult, thus the coarse rutile particles obtained γ-Na2TiO3 in which the atoms were orderly arrayed.

3.2.2. Effect of temperatures. As shown in Fig. 9a and b, α- and γ-Na2TiO3 changed into β-Na2TiO3 after heating at 800 °C, which means that β-Na2TiO3 is the thermodynamically stable phase at above 800 °C. There was an obvious endothermic peak at 815 °C in Fig. 6. Since the phase transition to β-Na2TiO3 could occur at 750 °C according to our previous study,21 815 °C may be a eutectic point of Na2CO3 + Na2TiO3. After slowly cooling the melt of β-Na2TiO3 at 1000 °C, the product remained as β-Na2TiO3, while quenching the melt yielded α-Na2TiO3 (see Fig. 9c and d).
image file: c6ra16984h-f9.tif
Fig. 9 XRD patterns of the product obtained by heating Na2TiO3 at different temperatures. (a) α-Na2TiO3, 800 °C; (b) γ-Na2TiO3, 800 °C; (c) β-Na2TiO3, 1000 °C, slow cooling; (d) β-Na2TiO3, 1000 °C, quenching. (e) Fine rutile powder mixed with a 50 wt% NaOH solution at ambient conditions, 500 °C.

Fig. 10 provides a summarized schematic illustration for the formation and transformation of α-, β-, and γ-Na2TiO3. γ- and β-Na2TiO3 are the thermodynamically stable phases of Na2TiO3 at approximately 500 °C and above 800 °C, respectively. α-Na2TiO3 can be prepared from the reaction of fine rutile particles and NaOH at 500 °C or quenching the melt of Na2TiO3 at 1000 °C quickly. It means that the fast kinetics of these processes prevents the Na2O–TiO2 system from reaching the thermodynamically stable state, and leads to the random distribution of Na and Ti atoms, thus forming the α-Na2TiO3 phase. It is reasonable to conclude that α-Na2TiO3 is formed far beyond the thermodynamically stable state. To verify the above deduction, TiO2 powder of D(50) = 1.5 μm was mixed evenly with a 50 wt% NaOH solution with a NaOH/TiO2 molar ratio of 2.2[thin space (1/6-em)]:[thin space (1/6-em)]1 in a nickel crucible and placed in a muffle furnace at 250 °C; the furnace was then heated to 500 °C for 90 min. Compared to the preparation of Fig. 2a, the addition of water accelerated the diffusion of Na2O, and allowed the O and Na atoms to reach the ordered array (thermodynamically stable state) more easily, which could break through the impediment of the reaction kinetics. In this case, γ-Na2TiO3 was obtained, as shown in Fig. 9e.


image file: c6ra16984h-f10.tif
Fig. 10 Schematic illustration for the formation and transformation of α-, β-, and γ-Na2TiO3.

3.3 Ion-exchange properties of α-, β-, and γ-Na2TiO3

Na2TiO3 remained unstable during the water washing because the Na+ in Na2TiO3 easily underwent ion-exchange with the H+ in water.17,22 This can be treated as the hydrolysis of Na2TiO3, which is described below:
 
Na2TiO3 + xH2O → Na(2−x)HxTiO3 + xNaOH (2)

As shown in Fig. 11a, the removal of Na+ increased with the increasing of water washing times at 25 °C. Na+ in β-Na2TiO3 exchanged slower than that in the α- and γ-Na2TiO3. The solubility of Ti(IV) was less than 0.001 mol L−1,51 which means that almost all Ti remained in the solids. Further study shows that the removal of Na+ after the 5th water washing at 65 °C were 84.5%, 79.5%, and 82.2% for α-, β-, and γ-Na2TiO3, respectively. The ion-exchange abilities can be arranged as follows: α-Na2TiO3 > γ-Na2TiO3 > β-Na2TiO3.


image file: c6ra16984h-f11.tif
Fig. 11 (a) The removal of Na+ of Na2TiO3 for different water washing times at 25 °C. XRD patterns of (b) α-, (c) β-, and (d) γ-Na2TiO3 after different water washing times and calcined at 250 °C.

XRD and Raman measurements were used to examine the samples after each water washing. Fig. 11 shows that α-, β-, and γ-Na2TiO3 converted amorphous phases after the 2nd, 6th, and 4th water washings at 25 °C, respectively. As shown in Fig. 11a, the exchange of Na+ was almost the same (about 70%) for the above washed sample (α-2, β-6, and γ-4) which was very interesting.

For α- and β-Na2TiO3, the XRD peaks of the washed samples in Fig. 11 were broadened and shifted, which may indicate that the lattice parameters changed because of a partial H substitution for Na. α-1 and β-5 exhibited similar peaks. α-1 had a typical FCC pattern, while β-5 was the same but with a rhombohedral distortion and a superstructure. However, the peaks of γ-Na2TiO3 became weaker and disappeared during the washing process, and no intermediate phase was found.

As shown in Fig. 12, the original Raman peaks of α-, β-, and γ-Na2TiO3 became weaker during the water washing. For α-Na2TiO3, the peak at 819 cm−1 moved to 884 and 888 cm−1 after the first and second water washings, and its strength rapidly decreased in the following washing. The Raman spectra of β-Na2TiO3 in the 5th and 6th water washings were very similar to that of α-Na2TiO3 in the first two water washings. The exchange of Na+ led to a symmetry reduction and a loosening of the structures, causing the blue shifts of these peaks and the distortions of the TiO6 units in α- and β-Na2TiO3. The peaks at 880–900 cm−1 corresponded to the very-short Ti–O bond between the terminal O atom and the central Ti atom of the distorted TiO6 octahedra.52 Our previous study showed that the layered structures formed during the ion-exchange process of α-Na2TiO3.17 β-Na2TiO3 was treated as the superstructure of α-Na2TiO3, and both had TiO6 units. Deduced from the XRD and Raman analysis, β-Na2TiO3 may break down at the beginning of the ion exchange and follow similar paths of ion-exchange with α-Na2TiO3. For γ-Na2TiO3, there were no new XRD or Raman peaks during the water washing. The rare TiO5 units caused it to follow completely different paths during the ion-exchange process, which will be studied in depth in our future study.


image file: c6ra16984h-f12.tif
Fig. 12 Raman spectra of the slurry obtained from washing (a) α-, (b) β-, and (c) γ-Na2TiO3 for different times at 25 °C. Numbers indicate the water washing times.

4. Conclusions

α-, β-, and γ-Na2TiO3 were prepared from rutile TiO2 and molten NaOH. Three models of β-Na2TiO3 with the space group of R[3 with combining macron], P[1 with combining macron], and P[3 with combining macron] were proposed, and the R[3 with combining macron] model was refined from the experimental data by Rietveld method. The structure of β-Na2TiO3 is a superstructure of α-Na2TiO3 and supposedly contains Ti6O19 clusters.

The crystal structures of Na2TiO3 were mainly determined by the particle sizes of rutile and the temperatures in the reaction of rutile and NaOH. Fine particle sizes of rutile favoured the formation of α-Na2TiO3. There was no crystal transformation between α- and γ-Na2TiO3 in the reactions below 600 °C. α-Na2TiO3 could be prepared from the reaction of fine rutile particles (D(50) < 25.8 μm) and molten NaOH at 500 °C or quenching the melt of Na2TiO3 at 1000 °C quickly. γ- and β-Na2TiO3 were the thermodynamically stable phases of Na2TiO3 at around 500 °C and above 800 °C, respectively. Thus, α-Na2TiO3 was formed far beyond the thermodynamically stable state.

The Na+ in α-Na2TiO3 was easier to exchange with H+ in water than that in β or γ phases. They all converted to amorphous phases after sufficient water washing. β-Na2TiO3 followed the similar paths of ion-exchange as α-Na2TiO3 in water washing process, which was different from that of γ-Na2TiO3.

Acknowledgements

This work was supported by the National Basic Research Program of China (Grant No. 2013CB632604), the National Science Foundation for Distinguished Young Scholars of China (Grant No. 51125018), the Key Research Program of the Chinese Academy of Sciences (Grant No. KGZD-EW-201-2), and the National Natural Science Foundation of China (Grant No. 51374191 and 51402303).

References

  1. J. Huang, Y. Cao, Z. Deng and H. Tong, J. Solid State Chem., 2011, 184, 712–719 CrossRef CAS.
  2. D. Nepak, RSC Adv., 2015, 5, 47740–47748 RSC.
  3. I. S. Grover, S. Singh and B. Pal, RSC Adv., 2014, 4, 51342–51348 RSC.
  4. X. Fu, D. Y. C. Leung and S. Chen, CrystEngComm, 2014, 16, 616–626 RSC.
  5. H. Zhang, L. Cao, W. Liu and G. Su, Appl. Surf. Sci., 2012, 259, 610–615 CrossRef CAS.
  6. H. Lu, RSC Adv., 2015, 5, 89777–89782 RSC.
  7. D. V. Bavykin, J. M. Friedrich and F. C. Walsh, Adv. Mater., 2006, 18, 2807–2824 CrossRef CAS.
  8. Y. Zhang, Z. Jiang, J. Huang, L. Y. Lim, W. Li, J. Deng, D. Gong, Y. Tang, Y. Lai and Z. Chen, RSC Adv., 2015, 5, 79479–79510 RSC.
  9. A. Rudola, K. Saravanan, S. Devaraj, H. Gong and P. Balaya, Chem. Commun., 2013, 49, 7451–7453 RSC.
  10. C. Y. Xu, J. Wu, P. Zhang, S. P. Hu, J. X. Cui, Z. Q. Wang, Y. D. Huang and L. Zhen, CrystEngComm, 2013, 15, 3448–3454 RSC.
  11. I. El-Naggar, E. Mowafy, I. Ali and H. Aly, Adsorption, 2002, 8, 225–234 CrossRef CAS.
  12. I. Ali, J. Radioanal. Nucl. Chem., 2004, 260, 149–157 CrossRef CAS.
  13. P. Sanchez-Camacho, I. C. Romero-Ibarra, Y. Duan and H. Pfeiffer, J. Phys. Chem. C, 2014, 118, 19822–19832 CAS.
  14. K. Zhao, Y. W. Wang, S. H. Tao and N. X. Feng, Adv. Mater. Res., 2013, 734, 2430–2433 CrossRef.
  15. D. Tang, D. Zhou, J. Zhou, P. Zhang, L. Zhang and Y. Xia, Hydrometallurgy, 2015, 157, 90–96 CrossRef CAS.
  16. F. Meng, T. Xue, Y. Liu, W. Wang and T. Qi, Hydrometallurgy, 2016, 161, 112–116 CrossRef CAS.
  17. Y. Liu, W. Zhao, W. Wang, X. Yang, J. Chu, T. Xue, T. Qi, J. Wu and C. Wang, J. Phys. Chem. Solids, 2012, 73, 402–406 CrossRef CAS.
  18. F. C. Meng, T. Y. Xue, Y. H. Liu, G. Z. Zhang and Q. I. Tao, Trans. Nonferrous Met. Soc. China, 2016, 26, 1696–1705 CrossRef CAS.
  19. A. J. Manhique, W. W. Focke and C. Madivate, Hydrometallurgy, 2011, 109, 230–236 CrossRef CAS.
  20. S. Middlemas, Z. Z. Fang and P. Fan, Hydrometallurgy, 2013, 131–132, 107–113 CrossRef CAS.
  21. F. Meng, Y. Liu, J. Chu, W. Wang and T. Qi, Can. J. Chem. Eng., 2014, 92, 1346–1352 CrossRef CAS.
  22. Y. Li, Y. Yang, M. Guo and M. Zhang, RSC Adv., 2015, 5, 13478–13487 RSC.
  23. F. Meng, D. Wang, T. Xue, W. Wang, Y. Liu and T. Qi, Asia-Pac. J. Chem. Eng., 2016, 11, 14–23 CrossRef CAS.
  24. Y. Liu, F. Meng, F. Fang, W. Wang, J. Chu and T. Qi, Dyes Pigm., 2016, 125, 384–391 CrossRef CAS.
  25. R. S. Roth, T. Negas, L. P. Cook and G. Smith, Phase Diagrams for Ceramists, The American Ceramic Society, 1981 Search PubMed.
  26. W. Antony Hill, A. R. Moon and G. Higginbotham, J. Am. Ceram. Soc., 1985, 68, C-266–C-267 CrossRef.
  27. C. E. Bamberger and G. M. Begun, J. Am. Ceram. Soc., 1987, 70, C-48–C-51 CrossRef.
  28. S. Mao, X. Ren and Z. Zhou, Chin. J. Struct. Chem., 2008, 27, 553–557 CAS.
  29. Y. Li, H. Y. Yu, Z. T. Zhang, M. Zhang and M. Guo, ISIJ Int., 2015, 55, 134–141 CrossRef CAS.
  30. V. B. Nalbandyan and I. L. Shukaev, Zh. Neorg. Khim., 1990, 35, 1085–1086 CAS.
  31. V. Nalbandyan, Russ. J. Inorg. Chem., 2000, 45, 1652–1658 Search PubMed.
  32. C. Xu, Y. Zhan, K. Hong and G. Wang, Solid State Commun., 2003, 126, 545–549 CrossRef CAS.
  33. D. S. Seo, J. K. Lee, E. G. Lee and H. Kim, Mater. Lett., 2001, 51, 115–119 CrossRef CAS.
  34. S. Y. Kim, K. K. Yu, I. S. Park, G. C. Jin, T. S. Bae and H. L. Min, Appl. Surf. Sci., 2014, 321, 412–419 CrossRef CAS.
  35. H. M. Kim, F. Miyaji, T. Kokubo and T. Nakamura, J. Biomed. Mater. Res., 1996, 32, 409–417 CrossRef CAS PubMed.
  36. A. C. Larsen and R. B. von Dreele, General Structure Analysis System (GSAS), Los Alamos National Laboratory Report LAUR, 2004, pp. 86–748 Search PubMed.
  37. B. H. Toby, J. Appl. Crystallogr., 2001, 34, 210–213 CrossRef CAS.
  38. S. K. Rout and S. Panigrahi, Indian J. Pure Appl. Phys., 2006, 44, 606–611 CAS.
  39. K. Momma and F. Izumi, J. Appl. Crystallogr., 2011, 44, 1272–1276 CrossRef CAS.
  40. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751–767 CrossRef.
  41. I. D. Brown, The Chemical Bond in Inorganic Chemistry. The Bond Valence Model, 2002 Search PubMed.
  42. I. D. Brown and D. Altermatt, Acta Crystallogr., Sect. A: Found. Crystallogr., 1985, 41, 244–247 CrossRef.
  43. Y. V. Kolen'ko, K. A. Kovnir, A. I. Gavrilov, A. V. Garshev, J. Frantti, O. I. Lebedev, B. R. Churagulov, G. Van Tendeloo and M. Yoshimura, J. Phys. Chem. B, 2006, 110, 4030–4038 CrossRef PubMed.
  44. B. Buchholcz, H. Haspel, Á. Kukovecz and Z. Kónya, CrystEngComm, 2014, 16, 7486–7492 RSC.
  45. B. C. Viana, O. P. Ferreira, A. G. Souza Filho, A. A. Hidalgo, J. Mendes Filho and O. L. Alves, Vib. Spectrosc., 2011, 55, 183–187 CrossRef CAS.
  46. X. Meng, D. Wang, J. Liu and S. Zhang, Mater. Res. Bull., 2004, 39, 2163–2170 CrossRef CAS.
  47. S. K. Barbar and M. Roy, J. Mol. Struct., 2012, 1024, 132–135 CrossRef CAS.
  48. K. Tsuzuku and M. Couzi, J. Mater. Sci., 2012, 47, 4481–4487 CrossRef CAS.
  49. G. W. Morey and J. S. Burlew, J. Phys. Chem., 1964, 68, 1706–1712 CrossRef CAS.
  50. M. Jansen, Angew. Chem., Int. Ed., 2002, 41, 3746–3766 CrossRef CAS.
  51. X. B. Li, S. W. Yu, N. K. Liu, Y. K. Chen, T. G. Qi, Q. S. Zhou, G. H. Liu and Z. H. Peng, Hydrometallurgy, 2014, 147, 73–78 CrossRef.
  52. D. Yang, H. Liu, L. Liu, S. Sarina, Z. Zheng and H. Zhu, Nanoscale, 2013, 5, 11011–11018 RSC.

Footnote

Electronic supplementary information (ESI) available: Procedures of obtaining the structures of β-Na2TiO3; atomic sites and bond valence calculations of β-Na2TiO3; morphologies of the prepared α-, β-, and γ-Na2TiO3. See DOI: 10.1039/c6ra16984h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.