Introduction

The crystal structures and functional properties of metal oxide materials are directly affected by the arrangement of their constituent MOn polyhedral building blocks1,2. Atomic-scale control over these arrangements is extremely important because 1: controlling the way to link MOn polyhedra together favors the creation of novel nanostructures, with a wide variety of particle sizes, crystal structures and morphologies3; 2: control of the local distortions and defect chemistry of the constituent MOn polyhedra can alter electrical properties such as local dipole moments and macroscopic polarization4; and 3: both the individual MOn configurations and their connectivities can profoundly affect the growth kinetics of metal oxides. Reliable control of the above factors enables the design of new solid-state materials and the associated development of novel practical applications.

TiO6-octahedral-based functional materials, e.g. several typical Ti-containing oxides (TiO2 and layer-structured titanates), exhibit versatile properties e.g. photovoltaic, electrochemical, catalytic and dielectric properties5,6,7,8. These functional oxides are built of the same fundamental TiO6 polyhedral building units (see e.g.Figure 1a) but connected differently9. The gaining of the necessary control over the local TiO6 octahedral arrangement is essential for property tailoring. Although a multitude of efforts have been devoted, the usual synthesis routes do not allow simultaneous control over structural/phase evolution, defect chemistry and resultant properties of the materials produced10. Otherwise, the synthetic strategies have to involve extremely complicated procedures and/or insurmountable challenges arising from using capping complexes that invariably either obscure the desired structural configurations or undermine surface reactivity. To date, only a few examples (mainly perovskites) refer to TiO6 octahedral tilting, rotating and/or distortions2, which are only fabricated in non-solution ways and what's more, these have to be subject to thermally/dynamically nonequilibrium states e.g. via sintering under high temperatures/pressures11, or based on external strains/stimulus12, or forming local intergrowth with non-stoichiometry13. As such, the control of TiO6 octahedral arrangements from equilibrium-state solution chemistries, i.e. the connectivities and configurations thereto on the atomic-scale level, remains unsuccessful. As a result, an investigation on TiO6 octahedral evolutions and the relationships between TiO6 octahedral arrangements and the resultant structures/properties is still missing.

Figure 1
figure 1

Representations of the connectivity and configurations of TiO6 octahedra and their control under solution reaction conditions allowing crystalline phase and morphology evolutions in TiO2 and layer-structured Lepidocrocite-type titanate.

(a) The connectivity and configurations of TiO6 octahedra in brookite, anatase and lepidocrocite-type titanate. The arrows denote the axial directions (e.g. x, y and z) of the crystal structure. The blue dashed lines and circles denote edge-sharing and vertex/corner-sharing connections, respectively. The white frames demonstrate the configurations in a TiO6 octahedron corresponding to respective structures, where the same (different) colors mean the same (different) length of Ti–O bonds. (b) TiO6 octahedral modification through pH controlling giving rise to the structural evolution (XRD patterns and SEM/TEM images). (c) SEM (left: RFe = 5%–20%) and TEM/HRTEM (right: RFe = 5% and 15%) images of as-prepared lepidocrocite-type titanate for the TiO6 octahedral modification controlled through the introduction of interfering (Fe3+) ions. Inset in c-right are the Fourier-transform diffraction patterns corresponding to the HRTEM.

There remains an urgent and ongoing need for new synthetic strategies enabling control over TiO6 octahedral configurations and connectivities. To this end, we herein report on a new, facile solution-chemistry approach with which the relationship between the synthesis variables and the resultant TiO6 octahedral networks is mapped out in terms of controlling simultaneously the pH values of solutions and interfering/blocking ions (Fe3+, Sc3+ and Sm3+). On the one hand, the interaction between H+/OH groups and TiO6 octahedral networks could respond to a strong negative/positive voltage and exhibit a large increase in electrical capacitance and subsequent variation in TiO6 octahedral arrangements14. The kinetics of nucleation/growth in solutions can thus be manipulated15, as theoretically predicted16, by the presence of H+/OH species since they influence the surface energies and surface tension, leading to variations in morphology and phase stability of TiO2. On the other hand, interfering/blocking ions could permit the local environment of the TiO6 octahedra to be altered17, introduce local chemical defects and change the local chemical potential energy of the particular TiO6 octahedra involved10,18. Likewise, introducing interfering ions could probably cause local distortion, even breaking the local symmetry of the TiO6 octahedra, giving rise to new local polarization states.

In the light of the two synthetic variables for tuning TiO6 octahedral configurations and connectivities described above, this paper presents the atomic-scale control of TiO6 octahedra in several typical titanium-containing oxides (e.g. TiO2 and layer-structured lepidocrocite-type titanate) from solutions and builds the correlation between TiO6 octahedral control and structures/properties.

Results

Like many other functional titanium-containing oxides, TiO2 (anatase, brookite and rutile) and titanates are all built out of the same basic structural unit, TiO6 octahedron. Figure 1a shows the arrangements of TiO6 octahedra in brookite, anatase and lepidocrocite-type titanate, showing distinctly different TiO6 octahedral configurations (i.e., different Ti–O bond lengths and angles) and connectivities (i.e., edge-sharing and vertex/corner-sharing connections). Depending on the local arrangements of TiO6 octahedra, the resultant oxides have different crystal structures. This can be considerably influenced by the reaction parameters, solution pH and the concentration of interfering ions, during the solution synthesis.

TiO6 octahedral connectivity and configurations controlled through pH tuning

The connectivity types as well as the local configurations of the individual TiO6 octahedra are strongly influenced by pH value (pH = 1 ~ 14) without interfering ions. Thereby, as a result of the TiO6 octahedral control under different pH conditions, the corresponding crystalline nanophases, morphologies/microstructures show considerable diversity (Figure 1b). The reaction condition of pH ≤ 6.0 usually gives anatase TiO2 nanophases (Figures 1b, S1) with edge-sharing connectivity of TiO6 octahedra (Figure 1a). The average crystallite sizes of the resultant compounds are estimated to be between 8.2 nm (pH = 1) and 21 nm (pH = 6) from the strongest diffraction line (101) for anatase using Scherrer equation, consistent with TEM result. It is interesting that rutile TiO2 seems able to exist only with pH range of 2.0–3.0 (S1). A high pH value, by contrast, promotes vertex-sharing connectivity of the TiO6 octahedra (Figure 1a) as evidenced by the appearance of brookite phase (Figure 1b) particularly with pH value of 10.5, which still gives a mixture of spindle-shaped brookite and square-like anatase phases. Pure brookite phase was achieved when pH value is varied between 12.0 and 14.0, as confirmed by the structural refinements (S2–S4). While, the resultant morphologies are highly pH-sensitive, i.e., flower-shaped assemblies with a dimension of ~2 μm at pH = 12.5 and well-dispersed spindle-shaped brookite particles with dimensions around 300 nm along the axial direction and 100 nm along the cross-sectional direction at pH = 13.5. When the pH value is beyond 14.0, a new secondary phase with the strongest diffraction at 2θ = 9.8° appears in addition to the dominant brookite phase (Figure 1b and S5), presenting two different morphologies: spindle-shaped brookite and rod-like layered titanates with scrolled TiO6 frameworks. This new phase is related to a type of sodium-containing titanate (EDS in S5) that has a corner/edge-sharing TiO6 octahedral connectivity (Figure 1a)19. Upon further increase of the alkalinity (OH), brookite is always the major phase unless extremely high alkali conditions such as the (OH) concentration 5 M reported for lepidocrocite-type titanate nanotubes (TNTs)20. Unfortunately, such an extremely high alkalinity would not be favorable to control the TiO6 octahedral units, nor the size, morphology and chemical compositions of the final products.

It is noted that only a few papers to date report obtaining pure brookite TiO2. Usually, it coexists easily with anatase, implying the difficulty in the formation of pure brookite TiO2 phase21,22. Even in this work, pure brookite TiO2 is limited to a relatively narrow pH value range. For this reason, the subsequent control over the configurations and connectivity of TiO6 octahedra using interfering ions will be based on high-purity brookite. As a matter of fact, all the nanophases/structures can be achieved by changing synthesis conditions from the starting brookite TiO2, so we note here that the controlling process can be generally considered as a process of TiO6 octahedral modification.

TiO6 octahedral connectivity and configurations controlled through the interfering ions

Dominant nanophase evolution

The introduction of interfering Fe3+ ions at pH = 12.5 leads to a nanophase transformation from brookite to anatase and/or lepidocrocite-type titanate, i.e. the TiO6 octahedra were modified. Pure-phase brookite was obtained when pH = 12.5 without interfering Fe3+ ion, while a small amount of Fe3+ ions, e.g.RFe = 5%, results in anatase TiO2 formation (S6 and S7), which is confirmed by the decreased intensity of the (121) to the (120) peak (see the detailed explanation in Figure S8), the appearance of a Raman peak at 515 cm−1 (S8) corresponding to anatase23, also the inhomogeneous morphologies observed (S7). The ratio of the brookite to the anatase phase is about 79: 21 in this specific case, estimated via XRD refinement (S9). When the RFe increases up to or beyond 10%, the TiO6 octahedral configuration and connectivity are mainly of lepidocrocite-type titanate form rather than anatase, suggesting that interfering Fe3+ ion favors the former. It is apparent that the interfering Fe3+ ion behavior differs by contrast with previous work where Fe3+ ions may either be stably doped into the anatase/rutile phase17,18, or a transformation from anatase to rutile was deduced10. Most importantly, the interfering ions seem to be dispatched into lepidocrocite-type titanate rather than remaining in brookite (S10).

As discussed above, interfering Fe3+ ions play an important, but still limited, role in the control of TiO6 octahedral configurations and connectivity for pH ≤ 12.5 because (i) both brookite and lepidocrocite-type titanate phases always coexist even at RFe = 20% (S6) and (ii) the morphologies were different due to the two-phase coexistence (S7). However, if the pH value is increased slightly up to 13.5, only a small amount of interfering Fe3+ ions (e.g., RFe = 5%) is required to achieve pure-phase sodium-containing titanate (clearly indicated by XRD (S11) and EDS (S12)) without a trace of either the brookite or anatase phase. Pure-phase lepidocrocite-type titanate can always be achieved over a broad Fe3+ amount range of RFe = 5–20% (S11). Note that without interfering Fe3+ ions, it is impossible to form the lepidocrocite-type titanate. Further increasing the pH value, e.g. beyond pH = 14, however, is not suitable for the control of TiO6 octahedral configurations and connectivity as the Fe3+ no longer behaves like an interfering ion as described above, instead forming iron-based oxide phases (S13).

Dominant microstructure evolution

Apart from the control of the connectivity and configurations of TiO6 octahedral units to achieve lepidocrocite-type titanate nanophases, the interfering Fe3+ ions also lead to macro/microscopic morphological and structural evolution. The SEM/TEM images (Figure 1c) demonstrate a distinct transformation in morphology from tubular structures to sheet-like nanoparticles with increasing RFe. Relatively low RFe (e.g. 5%) advances the formation and further rearrangement of distorted and scrolled TiO6 octahedral frameworks to form lepidocrocite-type titanate tubular structures. A typical tube possesses inner/outer diameter of around 50/120 nm and a length of up to several microns. These tubes are apparently different from those titanate nanotubes (TNTs) reported previously that were characterized by scrolling-up of 3–4 layered structures with inner/outer diameter of several/tens nanometers24. The resultant tubes in this work are highly crystallized with lattice spacings of 0.362 nm and 0.669 nm, corresponding to the (110) and (020) planes of the lepidocrocite-type titanates (Figure 1c-inset). Although TNTs have been well explored under concentrated alkali conditions (e.g., 10 M NaOH)24, we have not seen any report with respect to such titanate tubes to date, especially obtained with such a lower pH value.

After the intermediate stage where the lepidocrocite-type titanates show a mixture of short rod-like and round particles (e.g., RFe = 10%), the lepidocrocite-type titanate tubes are completely transformed into small-scale nanosheets when the RFe increases up to 15% (Figure 1c). These nanosheets are ~50 nm in the two-dimensional plane and highly crystallized with a d value of 0.362 nm that belongs to the (110) plane. Such a sheet-like morphology remains unchanged even when RFe reaches 20% (Figure 1c). We note that utilization of other interfering ions with different sizes (or charges) may greatly influence the resulting structures such as the cases of Sc3+ (S14) and Sm3+ (S15), 5% of which could form pure-phase lepidocrocite titanate with nanosheet shapes.

Microstructure identification

To examine the crystalline structures and locally TiO6 octahedral variations, spectroscopic techniques were used. For Raman, nine vibrations (Figure 2a) for all titanates were detected around 187, 279, 387, 442, 567, 663, 703, 785 and 906 cm−1, which match fairly well the structure with the orthorhombic layered lepidocrocite titanate (Figures 1a, 3b), a type of protonic acid 19,20. Nonetheless, the lengths of the Ti–O bonds and the symmetry of the corresponding octahedra are greatly changed, as indicated by apparent Raman shifts towards lower wavenumbers at 442 and 703 cm−1. The Raman band at 663 cm−1, with the same origin as the Ag mode of 703 cm−1, is associated with the bending of TiO6 octahedral layers during the formation of lepidocrocite-type titanates via scrolling of Ti–O networks25. With increasing RFe, this band gradually disappears, reflecting the transformation from titanate tubes to nanosheets. Systematic shift of Ti–O vibrations with increasing RFe is also reflected by FT-IR spectra (Figure 2b), e.g., a Ti–O vibration shifting from 679 to 657 cm−1. While, the vibration of interlayered Ti–O–Na bonds19 at 899 cm−1 does not show any shifts with increasing the RFe but only weakens gradually, indicating that a small amount of Fe3+ would be dispatched to the interlayer sites to occupy the original Na+ locations. EPR (Figure 2c, a detailed EPR is also see Figure S16) signal of geff = 4.3, a characteristic of isolated Fe3+ ions, is related to Fe3+ cations into an orthorhombic structure. Its decrease indicates that the diffusion of Fe3+ ions is progressed from their initial location at the oxide lattice towards the surface/interface/interlayer18. Consequently, gradual weakening of this signal also implies that a part of dopant Fe3+ ions was transported to the interlayered sites as those Fe3+ doped tetra-titanate bulks (K2Ti4O9), where Fe3+ ions occupy the sites of interlayered K+ ions26.

Figure 2
figure 2

Spectroscopic characterizations of lepidocrocite-type titanates with varying RFe.

(a) Raman spectra. (b) Furrier-Transition infrared (FT-IR) spectra. (c) Electron paramagnetic resonance (EPR) spectra. The curve from top to bottom denotes the spectrum of RFe = 5%, 10%, 15%, 20%, respectively.

Figure 3
figure 3

(a) Thermogravimetric (TG) and (b) Crystal structure of lepidocrocite-type titanate. a-inset is the differentiated TG (DTG) curve (black) of RFe = 15% and fitting results (colored). b-right shows the chemical coordination of TiO6 octahedron where the numbers denote the lengths (unit: Å) of Ti-O bonds.

For layered lepidocrocite titanate, a predominant characteristic is that interlayer ions are exchangeable with a variety of inorganic cations or organic groups27. One may thus assume that interlayer spaces can accommodate Na+ and/or H+ to compensate the charge neutralization. In terms of TG-DTG analysis (Figures 3a, S17), all the lepidocrocite-type titanates followed a similar three-fold dehydration process. For instance, the titanate formed at RFe = 15% loses 13 wt% of its total weight due to the dehydration of physisorbed water at 106°C, of chemisorbed water at 169°C and of interlayer protons at ~206°C, corresponding to weight losses of 5.4 wt%, 2.4 wt% and 5.2 wt%, respectively. Furthermore, ICP-AES measurement (Table S1) suggests the molar ratio of Fe to Ti coincides with their initial nominal ratios. With increasing RFe, the numbers of interlayer protons increased, accompanied by a slight decrease in the molar ratio of Na to (Fe+Ti) when RFe is beyond 15%. This, consistent with FT-IR and EPR, confirms that a small amount of Fe3+ dopants are probably dispatched into the interlayer Na+ sites. As a result, the structural components of the present lepidocrocite-type titanate can be expressed by the general formula , where the Na+ ions and/or protons (H+ or [H3O]+) or even Fe3+ cations co-occupy the inter-layer regions between the main framework of the TiO6 octahedra, as schematically shown in Figure 3b.

Mapping TiO6 octahedral connectivity and configurations under solutions for tunable nanophases and nanostructures

To achieve a comprehensive understanding of how to control the configurations and connectivities of TiO6 octahedra, experiments under a broad range of pH conditions in conjunction with a given RFe were performed, giving rise to tunable nanophases and nanostructures. When fixing RFe = 5%, solution reaction with a high alkalinity (pH = 14) leads to the occurrence of α-Fe2O3. While, low pH conditions (pH < 3.0) generally caused a significant loss of Fe3+ because of high solvency. A single-phase Fe3+ doped anatase would therefore form in the conditions of either weak alkalinity (pH = 9.0) or acidity (pH = 3.0). The pH values in between (S13) that, indeed, are the optimum pH conditions for growing Fe3+ doped anatase TiO2 nanocrystals28.

Once RFe reached 10%, a relatively high alkalinity leads to the coexistence of two (pH = 12.0) or three nano-phases (pH = 10.5) like anatase, titanate and brookite (S18). Nevertheless, the Fe-based oxide, e.g. α-Fe2O3, also appears even at a weak alkalinity of pH = 9.0, meaning that a high RFe and a relatively low pH value cause the formation of Fe-based oxides. A map outlining the nanophases and nanostructure evolution achieved is shown in Figure 4. This map provides a feasible and controllable route for the configurations and connectivity of TiO6 octahedral units controlled via pH values and Fe3+ dopant concentrations.

Figure 4
figure 4

Map of TiO6 octahedral modification through simultaneously controlling over the solution pH values and doping levels of interfering Fe3+ ions.

Effect of the structural evolutions on the giant dielectric responses

Giant dielectric permittivities (ε′) of ~104 were obtained at room temperature for all the lepidocrocite-type titanates (Figure 5a): ε′ decreases with increasing frequency and then turns flat in the higher frequency range. With increasing RFe, the lepidocrocite-type titanate shows a higher and frequency-independent ε′. Most strikingly, the titanate with RFe = 15% showed a ε′ of larger than 104 over a wide frequency range up to 105 Hz. Beyond RFe = 15%, the dielectric permittivity drops to a lower level with the fall threshold value at a relatively lower frequency. All titanates exhibit a similar Debye-like relaxation feature in ε″ curves29. This relaxation peak shifts towards the higher frequencies with increasing RFe, reaching a maximum value of 76.2 kHz for RFe = 15%. Afterwards, the peak, for the lepidocrocite-type titanate with RFe = 20%, shifts slightly towards lower frequency. The relaxation time (τ) is also estimated to be 395, 64, 4.2 and 9.4 μs for the lepidocrocite-type titanates of RFe = 5–20% by using the equation according to the relaxation peak position (fp) on ε″. These dielectric responses, with their fittings (capacitance |C| and Bode drawings in Figure S20) by taking grain boundary contributions into account, are related to structurally heterogeneous contributions from the grain and grain boundary regions30 as suggested by the impedance measurements (S19), which yields grain resistivities of 64 kΩ.cm, 23 kΩ.cm, 12 kΩ.cm and 14 kΩ.cm for the lepidocrocite-type titanates with RFe = 5–20%, while the resistivities for the grain boundaries are all beyond 10 MΩ, at least three orders of magnitude higher than that of the corresponding grains.

Figure 5
figure 5

Dielectric characterizations for Fe3+ doped lepidocrocite-type titanates with varying RFe.

(a) Frequency dependences of dielectric permittivities (real part ε′ and imaginary part ε″) measured at room temperature as well as the fitting curves (solid lines). (b) Arrhenius plots of grain conductivities (ln σg) versus measured temperatures (1/T). Solid lines denote the fitting results using Arrhenius law , where Ea is the activation energy, kB the Boltzmann constant, σ0 a constant related to the density of charge carriers and T the absolute temperature. Fitting the plots yields the Ea for lepidocrocite-type titanates with RFe = 5–20%. (c) Frequency dependences of normalized peaks, tanδ/tanδmax and M″/M″max of lepidocrocite-type titanates with RFe = 5% at given temperatures.

To gain insight into the underlying polarization mechanism, temperature-dependent conductivities (ln σg) were also recorded. Figure 5b summarizes the variations in ln σgvs. 1/T over the measured temperature region, which all show a well-defined linear characteristic in Arrhenius law29. Fitting the ln σg plots gives the activation energies (Ea) of 0.81, 0.73, 0.57 and 0.64 eV for the lepidocrocite-type titanates with RFe = 5–20%. Obviously, the lepidocrocite-type titanate of RFe = 15% possesses the smallest Ea, which can be associated with the optimum dielectric permittivity that will be addressed later. Further, M″/M″max combining with tanδ/tanδmax, an effective approach, was employed to determine the dynamic nature of polaronic particles. It is well-established that, the peak positions of the two curves for a delocalized or long-range transport must be overlapped31,32. Figure 5c clearly suggests a localized or short-range character for the polaronic particles as the peak position of the M″/M″max and tanδ/tanδmax curves is apparently not overlapped.

Discussion

Structural evolution by TiO6 octahedral modification

TiO2 (anatase, rutile and brookite) and lepidocrocite-type titanate are constructed of basic TiO6 octahedral structural units. Previous work suggests that TiO2 formation involves an intermediate titanate state followed by a deionization (e.g. Na+ and/or H+ ions) process22,33. Cations/anions during solution chemistry greatly affect surface charges and tensions, which in turn promote releasing Na+/H+ from titanate precursor to form TiO6 octahedra and further enable TiO6 octahedral reconstruction in terms of edge/vertex sharing to form brookite, anatase or rutile33,34,35. These TiO6 octahedral units themselves are structurally distorted and further become highly polarized when interacting with cations/anions, resulting in reconfigurations and reconnections in the solution reactions.

Generally, introducing heteroatom doping will affect the surrounding solution around heteroatom-containing TiO6 octahedra while resulting in defect states that could change the polarity of the TiO6 octahedra. The driving forces of the TiO6 octahedra control are closely related to the degree of tolerance of the defects for TiO6 octahedra in different crystalline polymorphs, which is intrinsically linked to the local environment of the Ti atoms, as well as the types of introduced interfering ions.

Brookite→Anatase

The TiO2 phase transformation from brookite to anatase promoted by the interfering Fe3+ ions can be ascribed to the differing structural arrangements of TiO6 octahedra between brookite and anatase. TiO6 octahedra in the anatase share edges along all three directions in a zigzag manner, while in the brookite, they become more complicated: TiO6 octahedra only share edges along the c direction, but alternately share vertices and edges along both a and b directions (Figures 1a, 6). When trivalent iron ions (Fe3+) are substitutionally doped into TiO2, defect states such as oxygen vacancies should be spontaneously introduced to maintain charge neutrality36. Meanwhile, the octahedra become highly distorted and polarized, which results in local fluctuations in structure and energy. Thus, sharing one more O atom (i.e., sharing two O atoms totally) with the neighboring TiO6 octahedron would be a better way to accomplish the defect compensation for the octahedra with Fe3+ situated, i.e., through edge-sharing connections. Once the edge-sharing connection for two octahedra happens, local fluctuations around the octahedra with Fe3+ situated in the solution would favor edge-sharing connections for more octahedra to develop into the three-dimensional anatase nanostructures. This statement complies well with the predictions made by density functional theory (DFT) using the generalized gradient approximation taking into account the Hubbard level (GGA+U)37,38, where convincing evidence that the presence of Fe3+ could strongly reduce the formation energy of defect states for anatase is provided. As a result, the TiO6 octahedra involving Fe3+ ions, in order to tolerate the defect states, will restack and turn themselves into anatase from brookite. An indirect evidence for DFT prediction is that interfering Fe3+ ions result in band-gap narrowing, in good agreement with experimental results (S21). The TiO6 octahedral modification with the phase evolution from brookite to anatase in the solution seems not obeying traditional crystal chemical theory where edge-sharing connection generally leads to cation-cation repulsions and then structural destabilization39. In fact, this is because the interfering Fe3+ ions play the key role in TiO6 octahedra modification, which allows a better tolerance of defect states in anatase. Perhaps this is why it is so difficult to synthesize high-quality brookite. We have not found any report to date to successfully introduce dopant ions into brookite structure.

Figure 6
figure 6

Schematic representations of the structural evolutions with varying the TiO6 octahedral connectivity and configuration.

(a) brookite TiO2. (b) lepidocrocite titanate. (c) anatase TiO2. Two rings denote the reversible circles of phase transformation and the corresponding directions are marked by arrows. The blue dashed ring and lines denote the vertex-sharing and edge-sharing connections of TiO6 octahedra, respectively.

Brookite→Titanate

The crystal structure of lepidocrocite-type titanate consists of two-dimensional corrugated skeleton layers of edge- and corner-sharing TiO6 octahedra (Figures 1a, 6c). The six Ti-O bonds in an octahedron can be classified into 3 pairs, since each pair in opposite directions has same length (Figures 1a, 3b). To this, the polarity (or symmetry) of the octahedra for lepidocrocite-type titanate should be lower (higher) than that for brookite but higher (lower) than that for anatase, since in brookite the lengths of six Ti–O bonds are all different, while in anatase four of them are short and two of them are long (Figure 1a)9. The substitutional Fe3+ ions, combined with those defects induced by interfering Fe3+ ions, further intensify the polarity and distortions of the TiO6 octahedra. When the concentrations of OH and Na+ ions in the solution are high enough, those TiO6 octahedra would need to slightly rotate and change the lengths of the Ti–O bonds to release the local strain, because: (i) lepidocrocite-type titanate has a relatively high similarity to brookite in TiO6 octahedral connectivity, which allows the control to be easily accomplished with only a small variation in energy; and (ii) OH and Na+ ions, as well as protons (H+ and/or [H3O]+) in the solution can bond to the TiO6 octahedra to compensate the charge neutrality and minimize the lattice relaxation, further stabilizing the octahedral skeletons. For lepidocrocite-type titanate, interlayer confined ions (Na+ ions and protons) are weakly bonded to the Ti-O skeletons, which are thus sufficiently flexible to be even altered and replaced27. When heterovalent Fe3+ ions were introduced, layer-structured titanate can tune the interlayer spacing to compensate for both the defect states and host-guest mismatching that are induced by the substitution of Fe3+ for Ti4+, accompanied with the accommodation of more Na+ ions and protons (H+ and/or [H3O]+) into the interlayers, for charge balance and TiO6 octahedral skeleton stabilization. This is why the contents of Na+ ions and the hydrated groups are slightly increased with increasing RFe as indicated by ICP and TG. By contrast, for brookite TiO2, due to the absence of lattice Na+/H+ ions, charge neutrality can only be compensated by those defect states (e.g. vacancies), spontaneously leading to structural instability for the Ti–O networks and thus recombination of TiO6 octahedra10,40.

Nanotube→Nanosheet (titanate)

Previous studies on several types of titanates indicate that a layered crystal structure is an important prerequisite for nanotube formation41. Differently, the formation of doped lepidocrocite-type titanate nanotubes here is a consequence of the interfering ions (Fe3+) that play a key role in TiO6 octahedral modification. Substitution of Ti4+ with the smaller-size and lower-valent Fe3+ gives rise to charge imbalance as well as internal strain in the TiO6 octahedra and even bending of Ti–O skeleton layers in order to compensate the elevated surface tension and free energy42,43. However, when RFe is relatively high, the TiO6 octahedra cannot accommodate so many Fe3+ counterparts and defect states. As a result, partial Fe3+ ions are situated at the interlayer sites of the original Na+ species. Since these Fe3+ ions are much smaller in size than that of the interlayered Na+ species, the spacing of the interlayer would slightly reduce, somewhat like the protonated sodium titanate after acid rinsing20. At this stage, the electrical repulsive interaction between two Ti–O skeleton layers would dramatically accelerate to produce relatively high Coulomb repulsive energy, which directly results in layered structure collapse as indicated by a theoretical study43. The occupational competition between the skeletal Fe3+ and the interlayered Fe3+ allows lepidocrocite-type titanate nanotubes to completely convert into the nanosheets.

Associated with the TiO6 octahedral control, the structurally mutual evolution between brookite, anatase and lepidocrocite-type titanate can be achieved through appropriately controlling the synthetic conditions of solution chemistries. The TiO6 octahedra in the as-prepared lepidocrocite-type titanates could be modified in the solution reactions and convert to anatase or brookite by deionizing, i.e., releasing Na+ (and/or doped Fe3+) species and protons (S22), which has been accepted as an important route for TiO2 synthesis.35 Conversely, Ti–O bonds in TiO2, as indicated from previous work, could react with the concentrated alkali, leading to lamellar fragments of Ti–O frameworks and finally reconstruct into titanates43,44. In summary, the mutual evolution between TiO2 and lepidocrocite-type titanate is the result of TiO6 octahedral modification, which can be well controlled in solution chemistries, as schematically demonstrated in Figure 6.

TiO6 octahedral modification for optimum giant dielectric response

Microstructure- and morphology-related physical and chemical properties of solids are of significant importance in fabricating functional nanostructures45. The dielectric properties of nano-based materials could be directly related to the microstructure of the solid, i.e. the nanoscale bulk and interfaces, where the solid can be defective or even polarized as a result of surface defect dipoles and/or surface charges, which are expected to be even larger when heteroatoms are involved. Earlier structural analysis identified that substitutional Fe3+ ions greatly influence the local environment and polarity of TiO6 octahedra thereby altering the macroscopic morphology. Since Fe3+ is slightly smaller in both size and valence than Ti4+, substitution of Fe3+ for Ti4+ would force local shrinkage of the lattice, resulting in local strain across a few atomic distances. Meanwhile, local disorder/distortion surrounding the Fe3+ will also appear due to dopant-host mismatching like many other doping cases46,47. That is, introducing Fe3+ could cause a local charge imbalance, leading to a polaronic-like distortion31. As a result, defect dipoles induced by e.g. ion-defect pairs, local defects or distortion would form to act as the available polaronic species, making a contribution to the high dielectric permittivity observed8,30,48. These defect dipoles, with short-range behaviour, are unable to catch up with the oscillation of the externally applied electric field, causing a quick decrease in ε′ at relatively high frequencies. Apparently, the number of these defect dipoles is proportional to RFe and its transport nature is determined by RFe, i.e., intrinsically linked to the TiO6 octahedral modification and microstructural nature.

It is well established that the transformation from lepidocrocite-type titanate nanotubes to nanosheets is a result of TiO6 octahedral control caused by the incorporation of interfering Fe3+ ions. Here, RFe has two impacts: one is an increase in the number of defect dipoles and the other is that hydrated groups (associated with protons, H+ and/or [H3O]+, etc.) and heavy Fe3+ instead of Na+ ions are partially dispatched into the interlayers. Our previous study indicates that inter-layered Na+ ions in sodium titanate nanotubes can easily migrate to induce charges at the interfaces and the conductivity of sodium-rich lepidocrocite-type titanate is one or two orders of magnitude higher than that of the protonated lepidocrocite-type titanate with more hydrated groups44. Thus, inter-layered Na+ ions in Fe3+ doped lepidocrocite-type titanates would not be expected to hinder the polaronic transport. Comparatively, hydrated groups are strongly bonded to TiO6 octahedral skeleton layers and meanwhile Fe3+ is too heavy to migrate, both of which thus act as a giant barrier to obstruct the polaronic transport49. Consequently, the competition of the above two aspects gives rise to the shortest τ, smallest grain σg and Ea, responsible for the optimum dielectric property. It should be noted that the giant dielectric response is not related to those space charges induced by the electronic hopping from possibly delocalized 3d1 of Ti3+ as there is no trace of Ti3+ detected by EPR (Figure 2c) and XPS (S23).

The primary goal of this work is to establish the structure/property correlation via atomic-scale control over the connectivity and configurations of TiO6 octahedra through solution chemistry. Further, when taking into consideration the fact that nearly all of the metal-oxide materials obtained are constructed by metal-oxygen polyhedral (MOn) units, one can envisage atomic-scale reaction mechanisms for synthesizing metal-oxide nanostructures, which is expected to remain an important and exciting subject for future studies, a judicious choice that enables simple bottom-up fabrication of structurally more complex metal-oxides or multifunctional materials and devices.

Methods

Sample Preparation

The synthesis of titanium-based oxide nanostructures was performed by solution chemistry using the following procedure: 0.04 mol TiOSO4 (Ourchem, 99%) was dissolved into 100 mL deionized water to form solution (I). 0.025 mol NaOH (Sinoreagent, 99.9%) was dissolved into 100 mL deionized water to form solution (II). Subsequently, solution (I) was slowly added into solution (II) while stirring. The resultant white suspension was thoroughly filtered and then dispersed in water to form a precursor solution through magnetic-ultrasonic stirring for 12 h. 2 M NaOH or H2SO4 (Sinoreagent, 99.9%) solution was then added to adjust the pH value. The resultant precursor solutions were then sealed in Teflon-lined stainless steel autoclaves and reacted at 220°C for 48 h for precipitation and crystallisation. The reaction products were filtered, washed with distilled water and dried in air to get the final samples.

Using a similar synthesis procedure, interfering Fe3+ ions (also the Sc3+ and Sm3+ ions for investigating the effects from different sizes of interfering ions), were then incorporated into the titanium-based oxide nanostructures to control the configurations and connectivity of the TiO6 octahedral units in an attempt to control the microstructures and properties of the resultant Ti-based oxides. The Fe3+-doped titanium-based nanostructures were achieved via the dissolution of the appropriate relative amount of TiOSO4 and Fe2(SO4)3 (Sinoreagent, 99.9%) into 100 mL of deionized water to give a total mole amount of 0.04 with mole ratios RFe = Fe/(Fe + Ti) with values of 5%, 10%, 15% and 20%, respectively. The remaining synthesis procedures were the same as outlined above for the un-doped titanium-based nanostructures. The preparation procedure of titanium-based nanostructures with doping other interfering ions i.e. Sc3+ and Sm3+ is similar to the case of Fe3+.

Sample Characterizations

The chemical compositions of the resultant samples were determined by inductively coupled plasma-atomic emission spectrometry (ICP-AES). The phase purity and crystallinity of the samples were characterised by X-ray diffraction (XRD, Rigaku Dmax2500, Cu Kα radiation, λ = 0.15418 nm). Ni powders were chosen as an internal standard for the determination of the peak positions. The average crystallite size, D, was calculated via the Scherrer formula, D = 0.9λ/(βcosθ), where λ is the X-ray wavelength employed, θ is the diffraction angle and β is defined as the half width after subtracting instrumental broadening. Structural refinements on the XRD patterns were performed using the Rietveld method. The particle size, morphology and microstructures of the samples were characterized using field emission scanning electron microscopy (SEM, JEOL JSM-6700), transmission electron microscopy (TEM, JEOL JEM-2010). Energy-dispersive X-ray spectroscopy (EDS) data were collected on INCA Energy 450 EDXA system.

Optical diffuse reflectance spectra of the samples were collected at room temperature via a Lambda 900 UV-vis spectrometer. Raman spectra of the final products were obtained on a UV-vis Raman System (Renishaw 1000) with an excitation wavelength of 532 nm. Their infrared optical spectra were also collected by a Perkin-Elmer IR spectrophotometer. Thermal analysis of the samples was performed by thermogravimetric (TG) analysis (Netzsch STA449C). The valence and defect states were studied by X-ray photoelectron spectroscopy (XPS, ESCA-LAB MKII spectrometer) and electron paramagnetic resonance (EPR, Bruker BioSpin EPR) at a microwave frequency of 9.86 GHz.

Electrical and dielectric measurements

The samples were also hot-pressed at a pressure of 12 MPa at 200°C for 30 min to make pellets with a diameter/thickness of 12/1.5 mm. Both sides of the pellets were then coated with Ag-based conductive electrodes and dried under infrared radiation. Alternative current (AC) impedance measurements were carried out using a precision LCR meter (Agilent 4294A) over a frequency (f) range from 40 Hz to 5 MHz at selected temperatures to obtain the complex impedance , where Z′(ω) and Z″ (ω) are real and imaginary parts, respectively. The complex dielectric permittivity ε*(ω) was then extracted based on the relation , where ω(= 2πf) is the angular frequency and C0 = ε0S/L represents the vacuum capacitance of a capacitor with an area of S and a thickness of L. These impedance parameters were also modeled using the least-squares refinement program EQUIVCRT to obtain an equivalent circuit through which the equivalent grain and grain boundary conductivity can be deduced.