Guo-Hui Pan*ab,
Tomokatsu Hayakawab,
Masayuki Nogami*b,
Zhendong Haoa,
Xia Zhanga,
Xuesong Quc and
Jiahua Zhang*a
aState Key Laboratory of Luminescence and Applications, Changchun Institute of Optics, Fine Mechanics and Physics, Chinese Academy of Sciences, 3888 Dong Nanhu Road, Changchun 130033, China. E-mail: guohui.pan@aliyun.com; zhangjh@ciomp.ac.cn
bDepartment of Materials Science and Engineering, Nagoya Institute of Technology, Showa, Nagoya 466-8555, Japan. E-mail: mnogami@mtj.biglobe.ne.jp
cDepartment of Physics, Changchun Normal University, Changchun 130032, China
First published on 12th October 2015
A heterobimetallic complex, zinc titanium glycolate acetate hydrate (Zn2Ti3–GAH), tentatively formulated as Zn2Ti3(OCH2CH2O)4(OCH2CH2OH)5(CH3COO)3·2HOCH2CH2OH·H2O, was synthesized by a room-temperature homogeneous precipitation in ethylene glycol solution. Its chemical composition, crystal structure, morphology, growth mechanism and thermal behaviors were characterized in detail. The precipitated Zn2Ti3–GAH was of a highly crystalline monoclinic phase and porous microrod morphology. As the single source precursor (SSP), Zn2Ti3–GAH was transformed into different phases of zinc titanate via thermal decomposition. With the remaining shape of the microrods, cubic phases of Zn2Ti3O8 and rutile TiO2 (r-TiO2) supported hexagonal phases of ZnTiO3 (h-ZnTiO3) were obtained by calcination at 500 and 700 °C, respectively; while r-TiO2 supported Zn2TiO4 were yielded in the form of dispersed particles or chains at higher temperature (950 °C). Benefiting from the SSP route and the confinement in the specific microrod domains of precursors, the heterostructures of r-TiO2–ZnTiO3 and r-TiO2–Zn2TiO4 were formed during programmable calcination. The studies on photocatalysis by degrading methylene blue (MB) under ultraviolet (UV) irradiation indicated that the as-transformed zinc titanate exhibited enhanced activity. In particular, r-TiO2 supported h-ZnTiO3 displayed the photodegradation reaction rate constant of 0.00163 s−1, which was comparable to that of commercially available Degussa P25 TiO2. This probably related to the more effective charge separation in the r-TiO2–ZnTiO3 heterostructure packed in the microrods.
The coordination systems of heterometallic complexes are mainly stabilized by alkoxo-bridges, polyolates chelating ligands and oxo ligands between different metal atoms,4 which could be accordingly termed as heterometallic alkoxides, heterometallic polyolate-alkoxides and heterometallic polyoxometallate, respectively. Heterometallic alkoxides were investigated much earlier.1,2 Examples of titanium containing heterometallic alkoxides such as [M{Ti(OR)5}]n (M = Li, Na, K), [Mg{Ti2(OEt)8Cl}(μ-Cl)]2, Cd{Ti2(OiPr)9}I, [{Cd(OiPr)3}Ba{Ti2(OiPr)9}]2 and M2Ti2(OR)x(acac)y (M = Mg, CoII, Ni) were synthesized through reactions involving union of two different metal alkoxides or reactions of titanium alkoxides Ti(OR)4 (R = alkyl) and alkali metal alkoxotitanate ligands with metal halides or acetylacetonates (acac);1,2,11 Some novel types of heterobimetallic polyolate-alkoxides derived from various polyols (glycols, di- and tri-ethanolamines (teaH3) with common formulas of MxTi(OGO)y(OiPr)z(HOGOH)n (M = Ce, Ta, Al, K, Na, Zr; G = CH2CH2, CMe2CMe2, CHMeCH2CMe2, CMe2CH2CH2CMe2) and MxTiy(A)z(OR)n (M = Mg, Ca, Sr, Ba, Al, Ta, Nb; A = diethanolaminate (2-), triethanolaminate (3-); R = iPr, tBu), etc., were generally obtained by the interactions of residual hydroxyl groups in the pre-formed homometallic derivatives (titanium polyolates) with alkoxides of other metals.3,4,9 More recently, a variety of heterobimetallic polyoxotitanate cage compounds with a general formulae of [TixOy(OR)zMnLm] (M = La, Ce, Nd, Eu, Ga, CoII, Zn, FeII, Cu, Ni, MnII, Mov, Crv, etc.; L = Cl, Br, I, HOiPr, McOH (Methacrylic acid), etc.) were synthesized by a well-controlled solvothermal method via a reaction of Ti(OR)4 or the pre-formed homometallic titanium-oxo-alkoxy cage with metal chlorides, hydrated metal acetates and hydrated metal sulfates.5–8,12–19 Synthesis at room temperature, however, was less reported.19 Organically-soluble crystals are frequently obtained during the solvothermal heating–cooling cycle or during a posteriori partial evaporation of the mother liquor.7 In particular, in the absence of the above mentioned stabilizing ligands U. Schubert et al. recently synthesized the Zn–Ti–POBC bimetallic complex by means of a bifunctional linker of p-carboxybenzaldehyde oxime (POBC-H) upon reacting Ti(OiPr)4 with Zn(POBC)2.20
In this study, we reported on a heterobimetallic glycolate–acetate complex, zinc titanium glycolate acetate hydrate (Zn2Ti3–GAH) tentatively formulated as Zn2Ti3(OCH2CH2O)4(OCH2CH2OH)5(CH3COO)3·2HOCH2CH2OH·H2O, through a facile room-temperature precipitation reaction in ethylene glycol (EG) solution, which started from ambient-stable titanium glycolate derivatives and zinc glycolate acetate species. Similar EG mediated routes were previously described by H. Fu et al. to synthesize Zn–Ti, Ni–Ti and Co–Ti heterobimetallic glycolate precursors, though full understanding of their chemical compositions and structures was lacked.21–23 Herein these concerns were detailed on the basis of various analysis techniques. The as-precipitated Zn2Ti3–GAH appeared as highly crystalline monoclinic phase and porous microrods morphology. It thermally decomposed into different crystallographic phase of ZnO–TiO2 system from phase-pure Zn2Ti3O8 (cubic, defect-spinel type, porous microrods), then to rutile TiO2 supported ZnTiO3 (hexagonal, ilmenite-type, porous microrods) and finally to rutile TiO2 supported Zn2TiO4 (cubic, spinel-type, microcrystallites). Of particular interest is ZnTiO3 as promising dielectric materials for capacitors or microwave devices,24–31 gas sensors,32 paint pigments,27,33 and novel luminescent host.24,34–36 Zn2TiO4 was attractive as regenerable sorbents for catalytic desulphurization and dehydrogenation of hot gas.37,38 Zn2Ti3O8 was recently demonstrated as the anode material for rechargeable lithium-ion battery and photocatalyst for water splitting.22,39 The work presented herein evaluated their photocatalytic performance against degradation of methylene blue (MB) under UV irradiations. r-TiO2 supported h-ZnTiO3 displayed higher photocatalytic performance probably due to the presence of r-TiO2–ZnTiO3 heterostructure packed in microrods.
To study the effects of Zn/Ti molar ratios of the precursors on the reaction process and precipitated products, additional synthesis with Zn:
Ti = 2
:
3 and 1
:
2 was performed under otherwise constant conditions.
To check the stability of Zn2Ti3–GAH against humidity, DMF and DMSO, 40 mg of the as-obtained sample was suspended in 15 mL of water, DMSO and DMF overnight, respectively; they were then collected for structure and morphology characterizations from the solution by centrifugation, washed with ethanol, and finally dried at 60 °C for 12 h in air. In the case of DMF, the soaking experiments were performed two times on the same powders.
Zn2Ti3–GAH was poorly soluble in water and common organic solvents.
FT-IR (KBr, 4000–400 cm−1): ν/cm−1 = 3425 br (νs(O–H) in EG and H2O), 2926 s (νas(CH2) C–CH2), 2856 s (νs(CH2) C–CH2), 1665m (δ(H–O–H)), 1570m (νas(COO−)), 1420m, (νs(COO−)), 1391w (δas(CH3)), 1340w (β(O–H)), 1200–1270w (νas(–C–C–O–)), 1116w (C–O–Ti), 1074 vs. (νs(C–O), δ(C–O–H)), 1034m (δ(C–O)), 914m (γ(CH2)), 885m (γ(CH2)), 624 s (νs(Ti–O)), 563 s (νs(Ti–O)), 478 s (νs(Ti–O)).
Solid-state 13C{1H} NMR (100.63 MHz, +25 °C, ppm): δ = 74.130, 63.910, 61.295. Solution 1H NMR (500.13 MHz, [D6] DMSO, +25 °C, ppm): δ = 1.059 (t, J = 5 Hz, CH3), 1.151 (s), 1.238(s, CH3), 1.250 (s, CH3), 2.016 (s, CCH3), 2.524 (t, J = 2 Hz, CH3), 3.256 (q, J = 5.5 Hz, OCH2), 3.315 (s, OCH2), 3.339 (s, OCH2 and H2O), 3.392 (t, J = 2.5 Hz, OCH2), 3.523 (q, J = 5.5 Hz, OCH2), 4.452 (p, J = 5.5 Hz, OCH2); solution 13C NMR (125.77 MHz, [D6] DMSO, +25 °C, ppm): δ = 39.900, 40.066, 40.232, 40.399, 40.476, 40.567, 63.249 (OCH2).
MS (m/z): calcd for 1273.22 (C34H81O36Ti3Zn, [M + H]+), found: 1273.00; calcd for 1159.05 (C28H64O29Ti3Zn2Na, [M + Na]+), found: 1158.93; calcd for 1137.06 (C28H65O29Ti3Zn2, [M + H]+), found: 1136.95; calcd for 1113.30 (C30H81O32Ti2Zn, [M + H]+), found: 1113.00; calcd for 943.12 (C24H56O26Ti2ZnNa, [M + Na]+), found: 942.85; calcd for 921.14 (C24H57O26Ti2Zn, [M + H]+), found: 920.85; calcd for 807.07 (C20H48O20TiZn2Na, [M + Na]+), found: 806.90; calcd for 785.09 (C20H49O20TiZn2, [M + H]+), found: 784.90; calcd for 727.16 (C20H48O20Ti2Na, [M + Na]+), found: 726.93; calcd for 705.18 (C20H49O20Ti2, [M + H]+), found: 704.93.
Elemental analysis (%): calcd for C28H64O29Ti3Zn2: Zn 11.48, Ti 12.61, C 29.52, H 5.66; found: Zn 12.34, Ti, 12.10, C 27.29, H 5.86.
TGA/DTA: theoretical weight loss 64.8%, actual 66.7%. Thermal events (°C): 180w (endo), 214s (exo), 291s (exo), 370w (exo).
![]() | (1) |
![]() | (2) |
Zn(CH3COO)2·2H2O or anhydrous Zn(CH3COO)2 were found to produce crystalline Zn–EG–AC complex nanowires when refluxed in EG, which definitely comprised both EG species and acetate ions upon IR analysis.45,46 Alcoholate groups were considered to be formed through the alcoholysis of acetate ions with alcohol due to its intrinsically weak basicity when dissolving some hydrated acetate salts in polyols medium.47 Such a polyols (EG, glycerol, etc.) mediated process was employed to synthesize many other crystalline homometallic polyolate complexes (e.g. Sn(II), Pb(II), Co(II), Mn(II)) powders by reacting a metallic salt in appropriate polyalcohols instead of monoalcohols under heating.47–49 Herein at room temperature and in EG Zn(CH3COO)2·2H2O was suggested to undergo pre-alcoholysis to de-coordinate acetic acid, producing zinc glycolate acetate species. They then interacted with titanium glycolate precursors to yield the Zn2Ti3–GAH, which precipitated from the reaction medium owing to low solubility.
Herein the Zn/Ti stoichiometric ratio (2:
3) of Zn2Ti3–GAH is lower than that used in the starting materials (1
:
1), excess zinc acetate was left in the solution and discarded during the subsequent washing. Additionally, it was experimentally found that adopting the Zn/Ti molar ratio of 2
:
3 or 2
:
1 while keeping the constant concentration of titanium source displayed no effects on the chemical composition and crystal structure of the resulting Zn2Ti3–GAH. It means that controlled stoichiometric incorporation of Zn(II) and acetate ligand into the titanium glycolates is impossible. However, the morphology, precipitation process and production yield were affected. The appearance of milky colloids was delayed and the production yield was decreased at lower Zn/Ti ratio. With the stoichiometric ratio (2
:
3) of Zn/Ti, the first visible sign of the turbidity appeared in the clear solution ∼1.5 h after the addition of zinc source while the yield was decreased to ∼38% with respect to [Ti(OC4H9)4] supplied.
In the FT-IR spectrum of Zn2Ti3–GAH complex (see Fig. S1†) the strong and broad band associated with O–H stretching mode in the range of 3050–3500 cm−1 and the band at ∼1665 cm−1 associated with the H–O–H bending mode evidenced the hydrate for Zn2Ti3–GAH complex. In a medium wavenumber range, the indicative absorption bands centered at ∼1570 and 1420 cm−1 show that the acetate ligands (CH3COO−) were definitely involved in the Zn2Ti3–GAH complex. They were suggested to coordinate to the titanium atoms in a chelating coordination fashion given the relatively small difference between asymmetric and symmetric stretching vibrations (Δν = νa(COO−) − νs(COO−) = 150 cm−1).17,50 In a lower wavenumber range, absorption bands centered at 1074 cm−1 (C–O stretching and C–O–H bending) and 885 and 914 cm−1 (CH2 rocking vibration) were attributed to the modes of EG molecules and glycolate ligands. In the frequency range much lower than 700 cm−1, the sharp absorption bands are assigned to stretching and bending vibrations of Ti–O bonds.51
The as-precipitated Zn2Ti3–GAH was highly crystalline (see Fig. S2†). However, such a XRD pattern cannot be indexed in the Inorganic Crystal Structure Database (ICSD) or International Centre for Diffraction Data (ICDD). A preliminary searching for peaks followed by analysis using McMasille program reveals a primitive monoclinic system (figures of merit M (20) = 64.85). The unit cell parameters were primarily determined to be a = 7.4564 Å, b = 9.4574 Å, c = 7.5544 Å, α = 90°, β = 120.093°, γ = 90°, volume = 460.917 Å3, which is extremely close to a hexagonal unit cell. The best space group estimated by Checkgroup is P21/n. No crystals suitable for single crystal X-ray diffraction analysis could be grown for Zn2Ti3–GAH. Presently it is beyond our ability to predict precisely the connectivity of glycolate and acetate ligands, and then to refine the molecule and crystal structures of Zn2Ti3–GAH. However, previous studies on the homometallic glycolates (Ti(IV), In(III), etc.) nanowires prepared by refluxing glycols and metal alkoxides or metal salts suggested that these complexes shared a chain-like structure.41 Based on the similarities between our reaction and those reported in the literature, and the observation of a wire-like morphology, we believe that Zn2Ti3–GAH could also bear similar chains. A schematic representation of suggested structural arrangement of Zn2Ti3–GAH complex was shown in Scheme 1.
SEM image of Zn2Ti3–GAH clearly shows that the crude particles (Fig. 1A) are mainly monodisperse microrods with the average diameter of ∼0.5–1.5 μm and length of ∼4.0–6.0 μm. Higher-magnification of TEM image on single microrod (Fig. 1B) depicts that it looks like micromesh with varying contrast (lower or higher) spanning the whole microrod. The presence of interior cavity implies that these microrods are of hierarchical structures assembled or stacked by smaller building blocks (∼10–20 nm in size). Nitrogen sorption experiments (Fig. S3A†) showed a broad distribution of pore size from micropore to large pore, with sharp peaks at ∼2.4 and 3.8 nm. The BET surface area was ∼21.28 m2 g−1. Contrary to power XRD analysis, high-resolution TEM (Fig. 1C) and SAED (see inset of Fig. 1B) on single microrod did not reveal the high crystallinity since no lattice fringes and diffraction rings or spots appeared under extensive observations. These are mostly likely due to the structural destruction by high-energy electron beam. The TEM-EDS analysis (Fig. S4†) revealed a Zn/Ti molar ratio near ∼2:
3, which agree well with the elemental analysis. In contrast, regular microrods were also observed at Zn/Ti ratio lower than 2
:
1, but many more irregular particles along with microrods appeared when adopting the Zn/Ti ratio lower than 1
:
1.
FT-IR spectrum of Zn2Ti3–GAH definitely indicates the presence of glycolate and acetate ligands, however, the solid-state CPMAS 13C NMR spectrum of Zn2Ti3–GAH (see Fig. S5†) only displayed three resonance signals at δ ∼ 70 ppm for methylene [–OCH2–] group (two stronger signals at δ 74.130, 61.295 ppm and one weak 63.910 ppm), no resonances around δ ∼ 25 (CH3 region) and ∼180 ppm (COO region) related to acetate groups were detected mostly likely due to its low content.52 The signal at 74.130 ppm was assigned to chelated glycolate ligand, while the signal at 61.295 and 63.910 ppm were assigned to neutral glycols molecules.44 In solution only one resonance of 13C of free glycols was detected at ∼63.249 ppm (Fig. S6F†). Herein crystallization of EG solvent in the microstructure of Zn2Ti3–GAH and then the interactions of coordination and hydrogen bonding, were considered to shield the principle resonance to ∼61.295 ppm. In contrast, solution 1H and 13C NMR spectra (Fig. S6†) exhibit a number of peaks fallen in the region of methyl and methylene groups due to the dissociation of Zn2Ti3–GAH and/or further reactions in deuterated DMSO. Indeed, immersing Zn2Ti3–GAH in DMSO changed the crystallinity and morphology greatly (see Fig. S7B and S8C and D†). The materials became amorphous after soaking overnight; many separated small nanoparticles (∼50 nm) were released along with the remaining microrods. The DMSO solvent damaged the Zn2Ti3–GAH mostly likely starting from the extraction of neutral molecules (EG and H2O) in the microstructures.
Similar to many other crystalline homometallic glycolate powders,40,41,47–49 the solubility of Zn2Ti3–GAH was also poor in common organic solvents, however, molecule ion peak of which in MS (see Fig. 2B) was detected by using DMF at m/z 1136.95 (theoretical 1137.06), corresponding to [C28H65O29Ti3Zn2]+, and the experimental isotopic pattern (Fig. 2C) matches exactly with the simulated pattern (Fig. 2D). Considering the outstanding high stability of titanium glycolate toward moisture,10,42 the present as-synthesized complex was suggested to contain no oxo ligands inside. In addition, given the starting materials and solvent used, molecular ligands determined by FTIR, thermal evolution by TG-DTA discussed below as well as charge balance held in the molecular formula, Zn2Ti3–GAH was thus tentatively formulated as Zn2Ti3(OCH2CH2O)4(OCH2CH2OH)5(CH3COO)3·2HOCH2CH2OH·H2O. Additionally, MS also shows several molecule ion peaks of other species (see Fig. 2A and B and S9†). It was considered that a small amount of Zn2Ti3–GAH was destroyed and/or dissociated into many other complexes of lower nuclearities when dispersed in solvents for MS analysis. Similar to the case of DMSO, soaking Zn2Ti3–GAH in DMF overnight also changed the crystallinity and morphology but proceeded slowly. After one time soaking the diffraction peak intensity was dramatically reduced (Fig. S7C†), and then the materials became completely amorphous without any diffraction peaks after second immersion (Fig. S7D†); the surface of microrods became scraggly with a small amount of ∼50 nm nanoparticles (see Fig. S8E–H†). Additionally, after immersing Zn2Ti3–GAH in H2O overnight, as shown in Fig. S7E and S8I and J,† the materials also became amorphous; many microrods were retained but with cavities inside, some microrods were broken into smaller pieces. It is apparent that solvents destroyed the Zn2Ti3–GAH in different ways.
The thermal properties of Zn2Ti3–GAH were also examined by TG-DTA in air. As shown in Fig. 3, S10 and S11,† thermal behaviors could be mainly divided into four steps, namely, initial endothermic stage, subsequent two exothermic combustion steps, and a last exothermic process. Each process, especially the initial endothermic stage and subsequent combustions, resulted in a great mass loss. The total weight loss was ∼66.6% when heated to 1000 °C. The endothermic peak at ∼180 °C associated with a ∼39.0% mass loss was in agreement with the theoretical loss upon considering the departure of surface moisture, hydration water, lattice EG molecules and partially deprotonated glycolate groups (–OCH2CH2OH). EG is known to be a relatively stable organic compound with a boiling point of 197.6 °C.53 The intercalation of molecule EG therefore stabilizes the compound up to ∼180 °C. The first exotherm was associated with the combustion of totally deprotonated glycolate ligands [(OCH2CH2O)4], with a mass loss of ∼10.6% well comparable with the calculated value. The second sharp exotherm was attributed to the combustion of acetate ligand, with a weight loss (∼13.1%) slightly lower than the calculated value (∼15.5%) due to the progressive release. The last weaker exothermic process continued in a broad temperature range was related to the slow burnout of the organic residuals and crystallization behaviors of zinc titanate.
TG-FTIR was used to examine the gas species evolved during heating of Zn2Ti3–GAH. Fig. 4 displays a two-dimensional (2D) wavenumber versus temperature FTIR spectrum. At temperature below ∼52 °C, no vibration absorption could be detected, then some apparent absorption bands related to hydroxyl (O–H, 3650 cm−1), methylene (CH2, 2945, 2886 cm−1), carbonyl (CO, 1731 cm−1) and C–O groups (1051 cm−1) appeared, indicating the release of water, EG molecules and even some acetic species. In accordance with the endotherm in DTA, these bands became relatively stronger at ∼200 °C. The C–H and C–O modes were still detected during the first exotherm (∼210 °C) and but disappeared in the second one (∼305 °C). It means that glycolate species underwent insufficient combustion into H2O and CO2 during the first exotherm and were completely removed from the solid structure below ∼305 °C. The CO2 modes in the range of 2265–2400 cm−1 started to appear after ∼170 °C and nearly disappeared after the second combustion (∼400 °C). They were mainly associated with the oxidation process and accordingly displayed stronger absorptions during the combustion processes. The absorption band of carboxyl ligand retained up to 700 °C, also indicating the slow release of acetate species. These carbonyl containing species are most likely to be acetic acid, acetone, or acetic anhydride according to previous studies on Zn5(OH)8(CH3CO2)2·nH2O.52
In a typical synthesis, milky colloids appeared ∼1 h after the addition of zinc acetate solution. The growth of microrods in solution was studied by sampling at different periods of reaction time. As shown in Fig. S12A,† initially these colloids were mainly gel-like particles in morphology, which became less as the reaction proceeded. Rod-like particles appeared even after ∼1.5 h and dominated the products rapidly. The nucleation and growth proceeded continuously; some microrods grew preferentially into a relatively large size in a short time. As shown in Fig. S12B,† the dimension of the some rods reached a diameter of 0.4 μm and a length of 2.0 μm in ∼0.5 h. This was almost a limit of size in the following ∼3 h (see Fig. S12C–G).† Moreover, during growth process many microrods were characterized by ill-shape, rough side and end faces. It was therefore believed that the growth of microrods followed an aggregation/packing process of the as-grown small particles. They collided and then coalesced and/or self-assembled into microrods for surface energy minimization according to the chemical bonding theory of single crystal growth.54 In the reaction system, both polar solvent of EG and Zn2Ti3–GAH feature rich hydroxyl groups, which allow a variety of intermolecular forces such as hydrogen bonding and dipole–dipole electrostatic interaction, to promote the self-assembly into more dynamic and complex structures.55 In the following experiments, aging allowed for growing into much larger size, more regular shape and smoother surface via Oswald ripening.
As shown in Fig. 1, although the compositions as well as the structures of these zinc titanate are very different from the as-synthesized product, both of them held the shape of the Zn2Ti3–GAH until ∼700 °C. After calcination at 500 °C the diameter of microrods shrank to 0.5–1.0 μm, length to 2.0–3.0 μm (see Fig. 1D) due to the removal of organic moieties. TEM image of single microrod in Fig. 1E shows that the microrods are still porous but with a slightly larger pore size among the packed nanoparticles. The pore size distribution by BJH method upon nitrogen sorption analysis (Fig. S3B†) shows a sharp peak at ∼3.4 nm with a long tail in the regions of mesopore and large pore structures while the BET surface area increased to 34.28 m2 g−1 due to the removal of organic species. The SAED pattern (see the inset of Fig. 1E) on the microrod comprised some diffraction rings, indicating the polycrystallinity. These rings could be indexed to different crystal planes of single phase of cubic Zn2Ti3O8. High resolution TEM image (see Fig. 1F) exhibits some lattice fringes of the nanounits (∼5–10 nm) assembled in the microrods. The observed d-spacing of ∼0.25 nm is in good agreement with the lattice spacing of the (311) plane of cubic Zn2Ti3O8. At elevated temperature of 700 °C the resulting sample retained the rod shape, but the packed nanouints further grew into ∼100–400 nm in size and left much larger cavity in the microrods (see Fig. 1G and H). The pore size also shows a broad distribution from micropore to large pore with peaks at 1.7 and 30.1 nm upon N2 adsorption–desorption analysis (Fig. S3C†) while the BET surface area was dramatically decreased to 5.29 m2 g−1. In accordance with the XRD analysis that two phases of the h-ZnTiO3 and r-TiO2 were yielded under calcinations; herein it means that both phases occurred with each other in the microrod precursor, which could thus be termed as r-TiO2 supported h-ZnTiO3. Both h-ZnTiO3 (see Fig. 1I) and r-TiO2 nanoparticles (see Fig. 1K) were highly crystalline and imaged by high resolution TEM in single microrod where the interplanar spacing and corresponding crystal planes were indicated. Their FFT patterns (see Fig. 1J and L) characterized by regularly arranged spots indicate the single crystallinity. It is believed that a heterojunction of r-TiO2–h-ZnTiO3 formed in the particles-packed microrods. As shown in Fig. 1M and N, the Zn2TiO4 yielded upon thermal treatment at 950 °C did not retain the rod morphology, many bigger particles ∼0.4–1.0 μm in size were observed; some of them were grown into a chain. The porosity property accordingly disappeared. No typical isotherms with hysteresis loops were observed in nitrogen sorption experiment (Fig. S3D†). The BET surface area was only 0.02 m2 g−1. High resolution TEM on adjacent grains indicates the intergrowth of Zn2TiO4 (Fig. 1O) and r-TiO2 (Fig. 1Q) grains in one chain, forming a heterojunction structure (named as r-TiO2 supported Zn2TiO4). They are both of single crystallinity confirmed by SAED (Fig. 1P for Zn2TiO4; Fig. 1R for r-TiO2). The formation of r-TiO2 supported titanate obtained by post-treatment at elevated temperature was analogue to previous report on TiO2 supported Ce2Ti2O7.15
To date, a variety of routes have been developed to prepare zinc titanate (ZnxTiyOz) in the form of bulk solid ceramics, nano/microcrystalline powders, thin films, such as solid state reaction method,31,32,56 sol–gel processing,24–30 metallo-organic deposition (MOD) technique,38 radio frequency magnetron reactive sputtering.59 However, only a few studies have reported on the shape and size-controlled synthesis. Zinc titanate fibers were prepared by electrospinning.60 Zn2Ti3O8 nanowires were obtained via an ion-exchange reaction between the NaxH2−xTi3O7 nanotubes and a Zn2+ contained ammonia solution.58 Twinned Zn2TiO4 nanowires were synthesized using ZnO nanowires as a template or by an ordered assembly of nanobricks, respectively.61,62 One SSP route to zinc titanate films through sol–gel processing of Zn–Ti–POBC bimetallic complex was previously reported.20 The temperature-dependent structure evolution of ZnTiO3 by small angle X-ray scattering (SAXS) illustrated the advantage of SSP route to avoid the phase separation into regions with different compositions on the nanoscale level in the gel stage, compared to materials prepared from two individual precursors.20 In contrast, the SSP route presented herein benefits over the shape preservation of the precursors and in situ intergrowth of different phases (i.e. heterostructures) under suitable calcinations due to the confinement in the specific microrod domains.
Footnote |
† Electronic supplementary information (ESI) available: FTIR spectrum, XRD pattern, N2 adsorption–desorption isotherm at 77 K, TEM-EDS spectrum, CPMAS 13C NMR spectrum and solution 1H and 13C NMR spectra of zinc titanium glycolate acetate hydrate (Zn2Ti3–GAH) or other samples; XRD patterns and SEM images of Zn2Ti3–GAH after immersing in DMSO, DMF and H2O overnight; MS comparisons of the experimentally obtained isotopic distribution pattern with the calculated pattern of molecule ion peaks of other species detected after treating Zn2Ti3–GAH with DMF; the TG curve of Zn2Ti3–GAH in temperature range of 25–300 °C for showing the point of inflection between step 1 and 2 during thermal evolution; SEM images of products by sampling at different periods of reaction time when synthesizing Zn2Ti3–GAH; determination of the band gaps of Zn2Ti3–GAH and zinc titanate via thermal treatment at varied temperature. See DOI: 10.1039/c5ra18292a |
This journal is © The Royal Society of Chemistry 2015 |