Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Polyoxoniobates as molecular building blocks in thin films

Mark A. Rambaran , András Gorzsás , Michael Holmboe and C. André Ohlin *
Department of Chemistry, Faculty of Science and Technology, Umeå University, 907 36 Sweden. E-mail: andre.ohlin@umu.se

Received 14th September 2021 , Accepted 30th September 2021

First published on 6th October 2021


Abstract

Niobium oxide thin films have been prepared by spin-coating aqueous solutions of tetramethylammonium salts of the isostructural polyoxometalate clusters [Nb10O28]6−, [TiNb9O28]7− and [Ti2Nb8O28]8− onto silicon wafers, and annealing them. The [Nb10O28]6− cluster yields films of Nb2O5 in the orthorhombic and monoclinic crystal phases when annealed at 800 °C and 1000 °C, respectively, whereas the [TiNb9O28]7− and [Ti2Nb8O28]8− clusters yield the monoclinic crystal phases of Ti2Nb12O29 and TiNb2O7 (titanium–niobium oxides) in different ratios. We also demonstrate a protocol for depositing successive layers of metal oxide films. Finally, we explore factors affecting the roughness of the films.


Introduction

To sustain the fast-paced energy- and technologically driven society we live in, we need to develop new materials, and improve on those currently in use. Many electronic devices, from mobile phones to MRI machines, rely on niobium pentoxide (Nb2O5) and tantalum pentoxide (Ta2O5).1 The inertness of niobium and tantalum oxides is associated with good biocompatibility and corrosion resistance.2,3 Important applications where niobium, tantalum and Nb/Ta-titanium oxide films are used include orthopaedic/orthodontic devices,4,5 waveguides,6 electromagnetic interference shielding,7 and coatings for corrosion protection and biomedical devices.8–11 Due to the extensive use of these metal oxides and a lack of alternatives, niobium and tantalum are classified as technologically critical elements.

Key challenges lie in sourcing – leading to coltan, the mineral mined to produce niobium and tantalum, being termed a ‘blood mineral’12 – and implementing sustainable recycling protocols due to the inertness of niobium and tantalum oxides. Hence, better atom economy in usage, coupled with improved recovery methods, are needed.

There is also a need to develop lead-free piezoelectric materials to circumvent the use of lead zirconium titanate (PZT) piezoelectric devices. Sodium potassium niobate (KNN) and lithium niobate (LiNbO3) are considered suitable alternatives to PZT.13 However, the methods of synthesising KNN thin films may be energy intensive, or lack reproducibility and require the use of toxic precursors,14 and the ability to vary the alkali–metal content and type is very limited, which is of relevance to this study.

Translating recent advances in the understanding of polyoxoniobate and -tantalate chemistry15 into novel methods for making thin films therefore promises to allow better control over the composition of the films, better atom economy, and access to films with improved or novel properties.

Polyoxoniobates and -tantalates are examples of polyoxometalates (POMs),16–18 discrete anionic oxide clusters of group 5 and 6 metals in their highest oxidation states, and they are attractive as potential building blocks in the preparation of extended materials, such as thin films. In particular, the discrete nature of POMs means that their structures can be known with confidence, and their structures and composition can in many cases be tuned by targeted synthetic modifications.15,19–25 At a minimum, this means that the elemental composition of the film can be controlled to a high degree, with the added possibility of the POM structure being wholly or partially retained in the solid film, so that the relative locations of individual atoms can be controlled.

Many POMs are also highly soluble – the solubility of tetramethylammonium ([N(CH3)4]+; TMA) salts of the polyoxoniobates typically exceeds 1 g ml−1 in water, and are moderately soluble in methanol and ethanol, making them easy to handle and manipulate in the lab. Recently developed protocols for microwave synthesis offer short reaction times (15 to 60 minutes),26 capable of yielding tens of grams of pure product in one-pot reactions in the lab. In addition, organic counter ions, such as TMA, can be removed by thermal decomposition,27,28 something which is not possible with alkali ions. This allows for the production of films that can be freely doped with different ions and at different ratios. Finally, while there is a range of existing methods for making metal oxide films,29–32 not all scale well, or can be applied to a wide range of substrates. Sol–gel,29 sputter deposition30 or laser ablation methods31,32 may suffer from issues of reproducibility and/or cost.32 The use of polyoxometalates circumvent these limitations, while offering the possibility of heteroatom doped Nb2O5 or Ta2O5 to be fabricated through soft methods. Film preparation through spin-coating of benign, non-volatile, water-compatible precursors that can easily be made in bulk quantities, is thus a very appealing idea.

Anhydrous niobium pentoxide (Nb2O5) – the most common form of niobium oxide – exhibits temperature-dependent polymorphism and can be made from amorphous hydrous Nb2O5 (or niobic acid). Pseudohexagonal TT-Nb2O5 forms between 500–600 °C, orthorhombic ortho-Nb2O5 (or T-Nb2O5) is obtained at 600 °C–800 °C, tetragonal M-Nb2O5 forms at 850 °C–950 °C, and monoclinic mono-Nb2O5 (or H-Nb2O5) is seen at >950 °C.33–35ortho- and mono-Nb2O5 prevail in the literature, with TT- and M-Nb2O5 being considered as metastable forms.33–35

Anhydrous Ta2O5 also exhibits polymorphism. It exists in an orthorhombic crystal system as L-Ta2O5 or ortho-Ta2O5 (low temperature form) below 1360 °C and reversibly converts to the H-Ta2O5 phase (high temperature form) above 1360 °C.36–38 Other Ta2O5 polymorphs are accessible but require high pressures to promote their formation.39 As with niobium, there exists a hydrated amorphous form, tantalic acid. The reactivity of these amorphous oxides is much higher than that of their anhydrous counterparts,26 and have been demonstrated to be good precursors in polyoxoniobate and polyoxotantalate synthesis.15,22–24,40

Titanium–niobium oxides are also found as orthorhombic and monoclinic crystal systems. Crystalline Ti2Nb10O29 is dimorphic and exists in both the orthorhombic and monoclinic crystal systems,41 while the crystalline TiNb2O7 exists only in the monoclinic crystal system.42 Both can be made from hydrous Nb2O5 (niobic acid) and TiO2 directly at ≥1000 °C,41–44 niobium citrate and Ti(iPrO)4,45 or NbCl5 and Ti(iPrO)4 as sources of hydrous Nb2O5 and TiO2, respectively,46,47 These methods are limited due to the extensive heating time and temperature required to form the titanium–niobium oxides from solid Nb2O5 and TiO2. Notably, the titanium–niobium oxides are considered as prospective anode materials for lithium-ion batteries, to replace graphite or Li4Ti5O12 anodes currently used.48,49 Studies of the electrochemical properties of Ti2Nb10O29 have determined it has a high rate capability and diffusion coefficient, which facilitates Li+ and electron transport.47,50,51 The TiNb2O7 is of particular interest because of its energy density and lithiation properties.48

Decaniobate (Nb10) has been used as an aqueous precursor for depositing Nb2O5 and Nb2O5-ITO thin films on glass to elicit optical, electrochromic and electrochemical properties.52–54 These studies were limited to temperatures below 400 °C in the annealing process of Nb2O5 due the thermal properties of ITO. Chemical condensation of Nb2O5 using formic acid was also investigated at room temperature.54

A comparative study by Fullmer et al.27 of the Nb2O5 and Ta2O5 thin films deposited from aqueous solutions of hexaniobate (Nb6) and hexatantalate (Ta6), respectively, found that smooth Ta2O5 films always formed from Ta6 – regardless of annealing temperature – unlike that of the Nb2O5 film from Nb6. There was, in particular, a noticeable increase in the roughness of the Nb2O5 films formed at 600 °C and 800 °C, when compared to Ta2O5 films. This was attributed to a preference of either a lower activation energy of grain growth in Nb2O5, or the formation of proton-bridged Ta6 dimers vs. linear proton-bridged Nb6 chain-oligomers, with the latter serving as nucleation seeds.27

Mansergh et al.28 further probed the orthorhombic Ta2O5 thin films deposited from Ta6. Film thickness was determined to be directly proportional to the concentration of Ta6, while film thickness decreased at elevated temperatures. The latter decrease is associated with the removal of water, and decomposition of the TMA as the film condenses. The Ta2O5 films also become courser following heating above 700 °C due to grain growth, which coincides with a decrease in the refractive index of the film above 600 °C.28 More recently, Saez Cabezas et al.55 have developed an electrochemical method for acid-catalyzed condensation of Nb2O5 and Ta2O5 thin films, from Nb6 and Ta6, respectively, via the potentiostatic oxidation of water. They determined both Nb2O5 and Ta2O5 film thickness to increase with the applied potential and that film thickness increases rapidly during the first two minutes of the electrochemical deposition process, after which it plateaus. Finally, Fast et al.14 spin-coated multiple layers of potassium sodium niobate thin films from sodium/potassium hexaniobate, yielding homogeneous films. Hexaniobate is only stable at alkaline pH above 9–10, which limits the chemical compatibility with potential dopants. In addition, the use of alkali salts leads to the forced inclusion of alkali cations in the finished film, whether desirable or not.

In this study, we used solution deposition of polyoxoniobates to develop surfaces where the crystallinity, thickness and roughness can be tuned by changing the starting material, concentration, and annealing temperature. A multi-step protocol has been developed to ensure layered metal oxide thin films can be deposited through an iterative spin-coating and annealing process.

Experimental

Chemicals were used as received. [NC(H3)4]6[Nb10O28]·2H2O (Nb10), [NC(H3)4]7[TiNb9O28]·3H2O (TiNb9), [NC(H3)4]8[Ti2Nb8O28]·6H2O (Ti2Nb8), [NC(H3)4]8[Nb6O19]·2H2O (Nb6) and [NC(H3)4]8[Ta6O19] (Ta6) were synthesised using microwave irradiation according to published procedures.26,56 Aqueous solutions (0.2 M) of Nb10, TiNb9, Ti2Nb8, Nb6 and Ta6 were spin-coated onto silicon wafers (50 × 50 mm, p-type, 10–20 Ω cm, Siegert Wafer GmbH, Germany) pre-treated with piranha solution and atmospheric-pressure plasma. See ESI for details. In the typical experiment, 100 μl of polyoxometalate solution (0.2 M) was placed at the centre of the substrate, and spun at 2000 rpm for four minutes, before annealing in a furnace using a ramped temperature program (see ESI for details). Raman spectra of bulk solutions of polyoxoniobates and of annealed powders were collected using a 532 nm laser on an Anton-Paar Cora 5600 spectrometer. Raman spectra of spin-deposited films were collected using a Renishaw Qontor Raman microscope, using a 405 nm laser. Powder X-ray diffraction measurements were done on a PANalytical X'Pert3 diffractometer with a CuK[small alpha, Greek, macron] source. Ellipsometry measurements were done using an Alpha-SE Ellipsometer (J.A. Woollam Co. Inc.). Atomic force microscopy imaging was done with a Bruker BioScope Catalyst instrument operating in peak force mode in air. See ESI for additional details.

Results and discussion

We have explored the spin-coating of tetramethylammonium (TMA) salts of a series of polyoxoniobate and polyoxotantalate clusters – Nb6, Nb10, TiNb9, Ti2Nb8 and Ta6 (Fig. 1). These targets were chosen as the structures are known with great confidence, and their solution phase chemistries have been studied extensively.57–60 While Ta6 and Nb6 are stable at pH 9, Nb10 is only stable in a very narrow pH region, ca. 5–7.5,57 above which it dissociates to Nb6. The replacement of one niobium atom by a titanium atom in Nb10, however, extends this stability range to pH 6.5–12.0 in TiNb9,59 and for Ti2Nb8 no dissociation or speciation is evident even at pH 12.5.60 While the microwave syntheses of Nb10, Nb6, Ta6 and TiNb9 have been reported,26,56 for this work we also developed the microwave synthesis of Ti2Nb8. Although these molecules can be made through hydrothermal methods by heating overnight,15,23,40 microwave synthesis affords these clusters after irradiation for 30 minutes, and can easily be scaled up to yield 10–15 grammes per one-pot batch. Rapid production of large quantities of compound is an important factor in making these types of clusters in industrial thin film synthesis both attractive and viable.
image file: d1dt03116c-f1.tif
Fig. 1 The [Nb10O28]6− (a), [TiNb9O28]7− (b), [Ti2Nb8O28]8− (c), [Nb6O19]8− (d) and [Ta6O19]8− (e) polyoxometalate ions. Red, light blue, grey and dark blue spheres represent oxygen, niobium, titanium and tantalum atoms, respectively.

Silicon wafers were chosen as the substrate in spin-coating deposition experiments as they have high smoothness, good chemical inertness, and can tolerate the elevated temperatures needed for annealing. Silicon substrates were cleaned with piranha solution – a 3[thin space (1/6-em)]:[thin space (1/6-em)]1 solution of concentrated sulfuric acid and 30% hydrogen peroxide – before being treated with atmospheric pressure plasma in a plasma chamber. This plasma treatment was necessary for the spin-coated polyoxometalates to adhere to the surface of the substrate.

First, the effects of the annealing temperature and polyoxometalate concentration were investigated. Based on spectroscopic ellipsometry, there is a consistent decrease in film thickness with increased annealing temperature up to 800 °C. The most notable change in thickness, however, occurs when the films are heated between 200 °C–400 °C and corresponds to removal of water and decomposition of the TMA counter ion. As expected, the film thickness increases at higher polyoxometalate concentrations. This is illustrated for the Nb10 and Ta6 films in Fig. 2. At high polyoxometalate concentrations (0.5 M), the films are prone to spontaneous crystallisation after spin-coating and prior to annealing, leading to the formation of a transparent crystalline layer on the surface of the film. After annealing, this crystallisation is manifested in the presence of grain boundaries – especially at higher temperatures – that induce film coarseness and total internal reflection of incident light. The latter phenomenon limits the reproducibility of thickness measurements. Hence, 0.2 M polyoxometalate solutions were used throughout the study as this concentration yielded sufficiently thick films that could still be probed via spectroscopic ellipsometry.


image file: d1dt03116c-f2.tif
Fig. 2 Thickness of Nb2O5 (a) and Ta2O5 (b) films obtained from annealing Nb10 and Ta6, respectively, and determined by ellipsometry.

Knowing that a phase change in Nb2O5 occurs between 800 °C and 1000 °C, we spin-coated aqueous solutions (0.2 M) of all polyoxometalates (Nb10, TiNb9, Ti2Nb8, Nb6 and Ta6) in water onto silicon wafers, and annealed the films at 800 or 1000 °C in air. Thereafter, the films were characterised by Raman spectroscopy, SEM and AFM.

Raman spectra of Nb10 films showed that mainly ortho-Nb2O5 is obtained after annealing at 800 °C, as indicated by the broad peak at ca. 690 cm−1 (Fig. 3a). Annealing at 1000 °C, similarly yielded mono-Nb2O5 as indicated by the Nb[double bond, length as m-dash]O stretch at ca. 993 cm−1. The crystal form of the film can thus be tuned by the choice of annealing temperature.


image file: d1dt03116c-f3.tif
Fig. 3 (a) Raman spectra of [Nb10O28]6− films annealed at 800 °C and 1000 °C, (b) Raman spectra and (c) PXRD patterns of powdered [Nb10O28]6− annealed at 200–1000 °C and (d) ortho-Nb2O5 and mono-Nb2O5 polymorphs of Nb2O5. Red spheres represent oxygen atoms and light blue polyhedra niobium atoms. See ESI (Fig. S2-5) for Raman signals of powdered [Nb10O28]6− annealed at 200–400 °C.

Powders of pure Nb10 were also annealed, to see whether this yielded the same products as annealed films. The Raman spectra of annealed powders showed the same behaviour as the thin films, with ortho- and mono-Nb2O5 dominating at 800 and 1000 °C, respectively (Fig. 3b). There is no evidence of residual TMA counter ion in either powder or film, indicating that it has been fully combusted during annealing. Powder XRD further confirmed the assignment of the crystal phases at 800 and 1000 °C, as well as showing that powders annealed at 200 and 400 °C remained amorphous (Fig. 3c). This agrees with previous results that determined amorphous Nb2O5 crystallisation commences at temperatures >400 °C.61Fig. 3d depicts the difference in packing of NbO6 octahedra within Nb2O5 crystal lattice of ortho- and mono-Nb2O5. Films of Nb6 behaved broadly the same as films of Nb10, yielding ortho-Nb2O5 and mono-Nb2O5 at 800 °C and 1000 °C, respectively (Fig. S2-2).

Raman spectra of films of TiNb9 annealed at 800 and 1000 °C looked very similar to those obtained from Nb10. However, a narrow band at ca. 997 cm−1 was observed at both temperatures (Fig. 4a). The thin film spectra were the same as those obtained from annealed TiNb9 powder (Fig. 4b), which allowed us to determine the composition of the film using PXRD to be ca. 80% 1[thin space (1/6-em)]:[thin space (1/6-em)]1 orthorhombic and monoclinic Ti2Nb12O29 (Fig. 4c), as well as 20% 1[thin space (1/6-em)]:[thin space (1/6-em)]1 orthorhombic and monoclinic Nb2O5, at both 800 °C and 1000 °C. The source of the Nb2O5 is two-fold: the titanium[thin space (1/6-em)]:[thin space (1/6-em)]niobium (Ti[thin space (1/6-em)]:[thin space (1/6-em)]Nb) ratio in Ti2Nb12O29 is higher than in the starting material, and Nb6 is present as a minor impurity in the one-pot synthesis of TiNb9 (Fig. S7-4). The peak at ca. 997 cm−1 in the Raman spectrum was assigned to Nb[double bond, length as m-dash]O symmetric stretching and the broad band in the 600 cm−1 region is shifted in Ti2Nb12O29 relative to Nb2O5, owing to the substitution of some of the NbO6 octahedra by TiO6 (Fig. 4d).


image file: d1dt03116c-f4.tif
Fig. 4 (a) Raman spectra of [TiNb9O28]7− films annealed at 800 °C and 1000 °C, (b) Raman spectra and (c) PXRD patterns of powdered [TiNb9O28]7− annealed at selected temperatures and (d) the monoclinic Ti2Nb12O29 titanium–niobium oxide. Red spheres represent oxygen atoms, light blue and grey octahedra represent niobium and titanium atoms, respectively. See ESI (Fig. S2-5) for Raman signals of powdered [TiNb9O28]7− annealed at 200–400 °C.

Films of Ti2Nb8 annealed at 800 °C and 1000 °C also exhibited Raman spectra that were consistent with the spectra of annealed of Ti2Nb8 powders at the same temperatures (Fig. 5a and b). The Raman peak at ca. 1000 cm−1 is symptomatic of Nb[double bond, length as m-dash]O symmetric stretching in TiNb2O7 and is consistent with the monoclinic crystal system. Raman peaks at ca. 646 cm−1 and 543 cm−1 are similarly attributable to stretches of NbO6 octahedra and Nb–O–Nb bonds, respectively. PXRD confirmed the presence of monoclinic TiNb2O7 at 800 °C and 1000 °C (Fig. 5c). However, Rietveld refinement showed that the monoclinic TiNb2O7 obtained after annealing at 800 °C contained ca. 16% 1[thin space (1/6-em)]:[thin space (1/6-em)]1 orthorhombic and monoclinic Ti2Nb12O29, while ca. 34% 1[thin space (1/6-em)]:[thin space (1/6-em)]1 orthorhombic and monoclinic Ti2Nb12O29 were present after annealing at 1000 °C. The Ti2Nb12O29 likely stems from thermal decomposition of Ti2Nb8 during the annealing process and the Ti[thin space (1/6-em)]:[thin space (1/6-em)]Nb ratio of TiNb2O7 is higher than the starting material. Decomposition of Ti2Nb8 was confirmed by the observation that treatment of an aqueous solution of Ti2Nb8 at 160 °C for one hour yielded signals in the 17O NMR spectrum that corresponded to a very low concentration of TiNb9 (Fig. S7-5). Notably, the shifting of the NbO6 octahedra stretching to higher frequencies is similarly attributable to the presence of TiO6 octahedra (substituting for NbO6 octahedra) within the TiNb2O7 crystal lattice (Fig. 5d); much like Ti2Nb12O29.


image file: d1dt03116c-f5.tif
Fig. 5 (a) Raman spectra of [Ti2Nb8O28]8− films annealed at 800 °C and 1000 °C, (b) Raman spectra and (c) PXRD patterns of powdered [Ti2Nb8O28]8− annealed at selected temperatures and (d) the monoclinic TiNb2O7 titanium–niobium oxide. Red spheres represent oxygen atoms, light blue and grey octahedra represent niobium and titanium atoms, respectively. See ESI (Fig. S2-5) for Raman signals of powdered [Ti2Nb8O28]8− annealed at 200–400 °C.

In contrast, annealing spin-coated films of Ta6 did not yield films with Raman spectra that agreed with their annealed powder counterparts, mainly because Ta6 films annealed at 800 °C and 1000 °C were Raman silent (Fig. 6a; only the background signal from the silicon wafer is observed). The Raman spectrum of powder obtained from annealed Ta6 at 800 °C and 1000 °C showed that it consisted of ortho-Ta2O5 (Fig. 6b), an assignment which was confirmed by PXRD (Fig. 6c). The broad peak at 895 cm−1 originates from the stretching of Ta–O bonds with different bond strengths due to corner/edge-sharing in tantalum oxide polyhedra. ortho-Ta2O5 exhibits a Raman peak at 845 cm−1 that is attributable to Ta–O stretching of TaO6 and TaO7 (pentagonal bipyramidal) polyhedra. The Raman peaks at 620 cm−1 and 700 cm−1 are from stretches of TaO6 octahedra, with the stretching at 700 cm−1 arising from Ta–O bonds with different bond strengths due to corner/edge-sharing polyhedra. The structure of crystalline ortho-Ta2O5 (Fig. 6d) depicts the different TaO6 and TaO7 polyhedra present in the crystal lattice.


image file: d1dt03116c-f6.tif
Fig. 6 Raman spectra of [Ta6O19]8− films annealed at 800 °C and 1000 °C, (b) Raman spectra and (c) PXRD patterns of powdered [Ta6O19]8− annealed at selected temperatures and (d) the ortho-Ta2O5 polymorph of Ta2O5. Red spheres represent oxygen atoms and dark blue polyhedra represent tantalum atoms. See ESI (Fig. S2-5) for Raman signals of powdered [Ta6O19]8− annealed at 200–400 °C.

The lack of a Raman signal from the Ta6 films was attributed to attenuation specific to the tantalum oxide films. This was confirmed by depositing films of solutions of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 Nb6[thin space (1/6-em)]:[thin space (1/6-em)]Ta6, which yielded strongly attenuated spectra relative to pure Nb6, whereas powders of the same solution showed no such attenuation (Fig. S2-6).

Decomposition of the TMA counter ion in the annealed films was probed through energy-dispersive X-ray spectroscopy (EDS) (Fig. 7). The elemental composition of C after annealing at 400 °C decreases below 10%, and did not exceed 1% after annealing at 800–1000 °C. This correlates to rapid decomposition of the TMA counter ion at 200–400 °C and the continued decomposition of any carbon residues (char) at higher temperatures. There is a notable increase in the metal (Nb, Ti or Ta) composition of the annealed films at 200–400 °C, in agreement with decomposition of TMA, which later plateaus at 800–1000 °C.


image file: d1dt03116c-f7.tif
Fig. 7 Atomic composition of annealed Nb10 (a), TiNb9 (b), Ti2Nb8 (c), and Ta6 (d) films. See ESI (Table S5-1) for details of full atomic composition. The carbon content decreases with annealing temperature in all cases.

The Ti content of annealed Ti2Nb8 films is higher than the Ti composition of annealed TiNb9, which corroborates the higher Ti content of Ti2Nb8 films. A spatial distribution of Nb and Ti in TiNb9 and Ti2Nb8 films annealed at 1000 °C, obtained from EDS mapping, is used to illustrate this in Fig. 8. Notably, the distribution of the elements confirms that the films are homogeneous.


image file: d1dt03116c-f8.tif
Fig. 8 EDS map of the spatial distribution of Nb (a), Ti (b) and Nb/Ti (c) in TiNb9, and Nb (d), Ti (e) and Nb/Ti (f) in Ti2Nb8 films annealed at 1000 °C.

Multiple depositions of films

We also investigated repeated deposition of layers of polyoxometalates. Nb10 was deposited on silicon wafers that had been prepared by plasma treatment, and then annealed. At 800 °C, the film after a single deposition consisted of ortho-Nb2O5, had a thickness of 147 ± 5 nm according to ellipsometry. This film was again treated by ambient pressure plasma, and a new layer of Nb10 was spin-coated onto the first one. Pre-treatment of the film by plasma was necessary to ensure good adhesion between the layers. This process of repeated spin-coating and plasma treatment was repeated to deposit four successive layers of film, and we did this with all five polyoxometalates included in this study. After four layers of Nb10 had been spin-coated onto the silicon wafer, a final thickness of 235 ± 5 nm, as determined by SEM, was obtained when annealing at 800 °C (Fig. 9a). Notably, SEM showed that the film possessed voids with no obvious layering, and the films were determined to be smooth (roughness of ±24 nm) via AFM.
image file: d1dt03116c-f9.tif
Fig. 9 Cross-section images for annealing: Nb10 at 800 °C (a) and 1000 °C (b); TiNb9 at 800 °C (c) and 1000 (d); Ti2Nb8 at 800 °C (e) and 1000 (f) films after multiple depositions. Films annealed at 800 °C showed no layering, while the films annealed at 1000 °C exhibited two layers above the silicon substrate.

In contrast, when successive deposition was carried out with annealing at 1000 °C, a film of mono-Nb2O5 with a total thickness of 642 ± 46 nm was obtained (Fig. 9b). Here, two layers were observed, with a lower layer of thickness 199 ± 6 nm, and an upper layer with thickness 442 ± 44 nm. Voids were also apparent, but only in the upper layer and it had a roughness of ±55 nm. The presence of a distinct first layer in the mono-Nb2O5 film is indicative of successive layered deposition and subsequent stacking of the mono-Nb2O5 film. Though stacking of layers is not discernible for the ortho-Nb2O5 film, the larger thickness of this film (235 ± 5 nm), compared to the first layer of the ortho-Nb2O5 film (147 ± 5 nm), may be due to coalescence of the successively deposited layers of the ortho-Nb2O5 film. The increase in roughness as the films were annealed at 1000 °C is attributable to enhanced grain growth at higher temperatures, which occurs simultaneously with the phase change.

Layering of TiNb9 and Ti2Nb8 films was also studied (Fig. 9c–f) and the thickness and roughness are summarised in Table S4-1. After four successive depositions of TiNb9 and Ti2Nb8 followed by annealing at 800 °C, Ti2Nb12O29 and TiNb2O7 films were obtained, respectively, with no discernible layering, as with Nb10. Furthermore, as with Nb10, two layers were obtained at 1000 °C for both titanium–niobium oxide films and roughness increased for Ti2Nb12O29 (±35.3 nm) but not TiNb2O7 (±13.8 nm). For both Ti2Nb12O29 and TiNb2O7 films annealed at 1000 °C, the lower layers of the films were homogeneous and free of voids, while the voids were apparent only in the upper layer, like with Nb10.

Successive deposition of four layers of Nb6 and Ta6 films annealed at 800 °C also yielded ortho-Nb2O5 and ortho-Ta2O5 films, respectively, which did not exhibit layering. The ortho-Nb2O5 film had voids (Fig. 10a) while the ortho-Ta2O5 film was homogeneous throughout (Fig. 10b). Furthermore, the Nb6 film annealed at 1000 °C produced a layered mono-Nb2O5 film – similarly to Nb10 – with both layers of film being homogeneous. The Ta6 film annealed at 1000 °C also formed a layered ortho-Ta2O5 film that was smooth and free of voids in both layers. The tendency for Ta6 to form smooth, homogeneous Ta2O5 films is likely a consequence of its tendency to gel, instead of crystallising like Nb6, as observed with the removal of water during rotary evaporation.27


image file: d1dt03116c-f10.tif
Fig. 10 Cross-section images for annealing: Nb6 at 800 °C (a) and 1000 °C (b); Ta6 at 800 °C (c) and 1000 °C (d). The tantalate films are noticeably smoother and more homogeneous than the niobate films.

The ortho-Nb2O5 film possessed more voids than the mono-Nb2O5 film, when comparing the ortho-Nb2O5 film as a whole to the lower layer of the mono-Nb2O5 film. In the solid state, Si–O–Nb bond formation is encouraged at the silica–Nb2O5 interface at higher temperatures.62,63 Therefore, more Si–O–Nb bonds will form when annealing the Nb2O5 film at 1000 °C, compared to 800 °C. This contributes to film densification at higher temperatures. The homogeneity of the lower layer of mono-Nb2O5 film in contrast to the presence of voids in the ortho-Nb2O5 film and in the upper layer of the mono-Nb2O5 film, is thus likely due to poor densification of the latter.

The formation of a distinct first layer in Ti2Nb12O29 and TiNb2O7 is similarly likely dependent on Si–O–Nb bond formation. However, Si–O–Ti bonds are considered to contribute to the binding of the film to the substrate as complete incorporation of Ti into the silica network occurs at high temperatures (1000 °C).64 The prevalence of this bonding is inherently lower than that of the Si–O–Nb bonds, based on the partial occupancy of Ti atoms within the crystal lattice of Ti2Nb12O29 and TiNb2O7.41,42 The presence of the Ti atoms within the crystal lattice contributes to formation of Nb–O–Ti bonds, which also facilitate the adhesion of subsequent layers of the titanium–niobium oxides and supports film densification. Hence, the upper layers of the Ti2Nb12O29 and TiNb2O7 films annealed at 1000 °C possessed fewer voids than the analogous layer of mono-Nb2O5 film.

Conclusions

The controlled deposition of metal oxide thin films is possible via an iterative spin-coating and annealing process of POMs. This approach allows deposition of alkali-free, crystalline metal oxide thin films with tuneable thickness and roughness. Notably, the optical and energy storage properties of metal oxide films is dependent on their crystal structures. The TMA salts of Nb6, Ta6, Nb10, TiNb9 and Ti2Nb8 clusters allow the deposition of orthorhombic and monoclinic Nb2O5, Ta2O5, Ti2Nb12O29 and TiNb2O7 from aqueous solutions, depending on precursor and conditions. The substitution of a combustible organic counter cation, here tetramethylammonium, for the more common alkali–metal counter ions not only allows for the deposition of alkali-free metal oxide films, but also allows for increased flexibility in the tuning of the composition and properties of the films. Furthermore, the ability to deposit films with tuneable crystallinity is crucial to developing high quality, pristine metal oxide thin films for advancing nanotechnology and semiconductor development. We also showed that Raman spectroscopy is a powerful and convenient method of determining the phase purity of niobium pentoxide (orthorhombic or monoclinic).

The inclusion of heterometal atoms in polyoxometalates leads to different solid-state structures, even when the difference in substitution is limited to a single atom. The site-specific inclusion of Ti within the decametalate structure is exemplary of this and it can be inferred that the site-specific inclusion of other heterometals may elicit a similar effect. There is thus great potential for developing this technique to include other types of heterometals, such as those that have been shown to be catalytically active in different reactions. By including these in a chemically inert substrate such as niobium pentoxide, chemical processes may be adapted to take place in environments that would otherwise be too demanding.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

C. A. O thanks the Kempe Foundation (JCK-2029.1) and the Swedish Research Council (2018-07039) for generous support. M. A. R thanks the Department of Chemistry at Umeå University for a postgraduate fellowship. Prof. William H. Casey at UC Davis and Dr Sven Albrecht from TANIOBIS GmbH, are equally thanked for generous gifts of hydrous Nb2O5 and hydrous Ta2O5, respectively. We also thank Prof. Markus Broström for providing a furnace for annealing experiments, and Drs Hussein Kanbar and Cheng Choo Lee (Umeå Core Facility of Electron Microscopy) for assisting with PXRD and FIB-SEM/EDS measurements, respectively. Support towards the Vibrational Spectroscopy Core Facility by the Chemical Biological Centre and the Department of Chemistry at Umeå University is greatly acknowledged.

References

  1. Y. Pozdeev, Qual. Reliab. Eng. Int., 1998, 14, 79–82 CrossRef.
  2. D. Pradhan, A. W. Wren, S. T. Misture and N. P. Mellott, Mater. Sci. Eng., C, 2016, 58, 918–926 CrossRef CAS PubMed.
  3. N. Wang, H. Li, J. Wang, S. Chen, Y. Ma and Z. Zhang, ACS Appl. Mater. Interfaces, 2012, 4, 4516–4523 CrossRef CAS PubMed.
  4. J. Ge, F. Wang, Z. Xu, X. Shen, C. Gao, D. Wang, G. Hu, J. Gu, T. Tang and J. Wei, J. Mater. Chem. B, 2020, 8, 2618–2626 RSC.
  5. Y. Guo, K. Xie, W. Jiang, L. Wang, G. Li, S. Zhao, W. Wu and Y. Hao, ACS Biomater. Sci. Eng., 2019, 5, 1123–1133 CrossRef CAS PubMed.
  6. Q. Zhao, R. O. Behunin, P. T. Rakich, N. Chauhan, A. Isichenko, J. Wang, C. Hoyt, C. Fertig, M. h. Lin and D. J. Blumenthal, APL Photonics, 2020, 5(116103), 1–8 Search PubMed.
  7. N. Maruthi, M. Faisal, N. Raghavendra, B. P. Prasanna, S. R. Manohara and M. Revanasiddappa, Synth. Met., 2021, 275(116744), 1–9 Search PubMed.
  8. M. Dinu, L. Braic, S. C. Padmanabhan, M. A. Morris, I. Titorencu, V. Pruna, A. Parau, N. Romanchikova, L. F. Petrik and A. Vladescu, J. Mech. Behav. Biomed. Mater., 2020, 103, 103582 CrossRef CAS PubMed.
  9. Y. Ge, Y. Wang, J. Chen, Y. Zou, L. Guo, J. Ouyang, D. Jia and Y. Zhou, J. Alloys Compd., 2018, 767, 7–15 CrossRef CAS.
  10. W. Hu, J. Xu, X. Lu, D. Hu, H. Tao, P. Munroe and Z.-H. Xie, Appl. Surf. Sci., 2016, 368, 177–190 CrossRef CAS.
  11. J. Xu, X. k. Bao, T. Fu, Y. Lyu, P. Munroe and Z.-H. Xie, Ceram. Int., 2018, 44, 4660–4675 CrossRef CAS.
  12. J. W. Mantz, Global Studies Review, 2008, 4, 12–14 Search PubMed.
  13. T. Ibn-Mohammed, S. C. L. Koh, I. M. Reaney, A. Acquaye, D. Wang, S. Taylor and A. Genovese, Energy Environ. Sci., 2016, 9, 3495–3520 RSC.
  14. D. Fast, M. Clark, L. Fullmer, K. Grove, M. Nyman, B. Gibbons and M. Dolgos, Thin Solid Films, 2020, 710(138270), 1–6 Search PubMed.
  15. M. Nyman, L. J. Criscenti, F. Bonhomme, M. A. Rodriguez and R. T. Cygan, J. Solid State Chem., 2003, 176, 111–119 CrossRef CAS.
  16. M. Nyman, Dalton Trans., 2011, 40, 8049–8058 RSC.
  17. L. Zhang and Z. Chen, Int. J. Energy Res., 2020, 44, 3316–3346 CrossRef CAS.
  18. H.-Y. Zhao, Y.-Z. Li, J.-W. Zhao, L. Wang and G.-Y. Yang, Coord. Chem. Rev., 2021, 443(213966), 1–23 Search PubMed.
  19. M. Amiri, N. P. Martin, C. L. Feng, J. K. Lovio and M. Nyman, Angew. Chem., Int. Ed., 2021, 60, 12461–12466 CrossRef CAS PubMed.
  20. N. P. Martin and M. Nyman, Angew. Chem., Int. Ed., 2021, 60, 954–960 CrossRef CAS PubMed.
  21. N. P. Martin, E. Petrus, M. Segado, A. Arteaga, L. N. Zakharov, C. Bo and M. Nyman, Chem. – Eur. J., 2019, 25, 10580–10584 CrossRef CAS PubMed.
  22. C. A. Ohlin, E. M. Villa, J. C. Fettinger and W. H. Casey, Angew. Chem., Int. Ed., 2008, 47, 5634–5636 CrossRef CAS PubMed.
  23. C. A. Ohlin, E. M. Villa, J. C. Fettinger and W. H. Casey, Dalton Trans., 2009, 2677–2678,  10.1039/B900465C.
  24. J.-H. Son and W. H. Casey, Chem. – Eur. J., 2016, 22, 14155–14157 CrossRef CAS PubMed.
  25. J.-H. Son, J. Wang and W. H. Casey, Dalton Trans., 2014, 43, 17928–17933 RSC.
  26. M. A. Rambaran, M. Pascual-Borràs and C. A. Ohlin, Eur. J. Inorg. Chem., 2019, 3913–3918,  DOI:10.1002/ejic.201900750.
  27. L. B. Fullmer, R. H. Mansergh, L. N. Zakharov, D. A. Keszler and M. Nyman, Cryst. Growth Des., 2015, 15, 3885–3892 CrossRef CAS.
  28. R. H. Mansergh, L. B. Fullmer, D.-H. Park, M. Nyman and D. A. Keszler, Chem. Mater., 2016, 28, 1553–1558 CrossRef CAS.
  29. P. Amaravathy, S. Sowndarya, S. Sathyanarayanan and N. Rajendran, Surf. Coat. Technol., 2014, 244, 131–141 CrossRef CAS.
  30. V. K. Balla, S. Banerjee, S. Bose and A. Bandyopadhyay, Acta Biomater., 2010, 6, 2329–2334 CrossRef CAS PubMed.
  31. V. K. Balla, S. Bodhak, S. Bose and A. Bandyopadhyay, Acta Biomater., 2010, 6, 3349–3359 CrossRef CAS PubMed.
  32. M. Grobelny, M. Kalisz, M. Mazur, D. Wojcieszak, D. Kaczmarek, J. Domaradzki, M. Świniarski and P. Mazur, Thin Solid Films, 2016, 616, 64–72 CrossRef CAS.
  33. E. I. Ko and J. G. Weissman, Catal. Today, 1990, 8, 27–36 CrossRef CAS.
  34. R. A. Rani, A. S. Zoolfakar, A. P. O'Mullane, M. W. Austin and K. Kalantar-Zadeh, J. Mater. Chem. A, 2014, 2, 15683–15703 RSC.
  35. H. Schäfer, R. Gruehn and F. Schulte, Angew. Chem., Int. Ed Engl., 1966, 5, 40–52 CrossRef.
  36. S. Lagergren and A. Magnéli, Acta Chem. Scand., Ser. A, 1952, 6, 444–446 CrossRef CAS.
  37. N. C. Stephenson and R. S. Roth, Acta Crystallogr., Sect. B: Struct. Sci., 1971, 27, 1037–1044 CrossRef CAS.
  38. N. C. Stephenson and R. S. Roth, J. Solid State Chem., 1971, 3, 145–153 CrossRef CAS.
  39. I. P. Zibrov, V. P. Filonenko, M. Sundberg and P.-E. Werner, Acta Crystallogr., Sect. B: Struct. Sci., 2000, 56, 659–665 CrossRef CAS PubMed.
  40. C. A. Ohlin, E. M. Villa and W. H. Casey, Inorg. Chim. Acta, 2009, 362, 1391–1392 CrossRef CAS.
  41. A. D. Wadsley, Acta Crystallogr., 1961, 14, 664–670 CrossRef CAS.
  42. A. D. Wadsley, Acta Crystallogr., 1961, 14, 660–664 CrossRef CAS.
  43. R. B. V. Dreele, A. K. Cheetham and J. S. Anderson, Proc. R. Soc. London, Ser. A, 1974, 338, 311–326 Search PubMed.
  44. X. Wu, J. Miao, W. Han, Y.-S. Hu, D. Chen, J.-S. Lee, J. Kim and L. Chen, Electrochem. Commun., 2012, 25, 39–42 CrossRef CAS.
  45. N. G. Eror and U. Balachandran, J. Solid State Chem., 1982, 45, 276–279 CrossRef CAS.
  46. S. Deng, Z. Luo, Y. Liu, X. Lou, C. Lin, C. Yang, H. Zhao, P. Zheng, Z. Sun, J. Li, N. Wang and H. Wu, J. Power Sources, 2017, 362, 250–257 CrossRef CAS.
  47. S. Lou, X. Cheng, J. Gao, Q. Li, L. Wang, Y. Cao, Y. Ma, P. Zuo, Y. Gao, C. Du, H. Huo and G. Yin, Energy Storage Mater., 2018, 11, 57–66 CrossRef.
  48. K. J. Griffith, Y. Harada, S. Egusa, R. M. Ribas, R. S. Monteiro, R. B. Von Dreele, A. K. Cheetham, R. J. Cava, C. P. Grey and J. B. Goodenough, Chem. Mater., 2020, 33, 4–18 CrossRef.
  49. Y. Yang and J. Zhao, Adv. Sci., 2021, 8(2004855), 1–24 Search PubMed.
  50. D. Pham-Cong, J. Kim, V. T. Tran, S. J. Kim, S.-Y. Jeong, J.-H. Choi and C. R. Cho, Electrochim. Acta, 2017, 236, 451–459 CrossRef CAS.
  51. C. Yang, S. Yu, Y. Ma, C. Lin, Z. Xu, H. Zhao, S. Wu, P. Zheng, Z.-Z. Zhu, J. Li and N. Wang, J. Power Sources, 2017, 360, 470–479 CrossRef CAS.
  52. A. Llordés, G. Garcia, J. Gazquez and D. J. Milliron, Nature, 2013, 500, 323–326 CrossRef PubMed.
  53. A. Llordés, A. T. Hammack, R. Buonsanti, R. Tangirala, S. Aloni, B. A. Helms and D. J. Milliron, J. Mater. Chem., 2011, 21, 11631–11638 RSC.
  54. A. Llordés, Y. Wang, A. Fernandez-Martinez, P. Xiao, T. Lee, A. Poulain, O. Zandi, C. A. Saez Cabezas, G. Henkelman and D. J. Milliron, Nat. Mater., 2016, 15, 1267–1273 CrossRef PubMed.
  55. C. A. Saez Cabezas, K. Miller, S. Heo, A. Dolocan, G. LeBlanc and D. J. Milliron, Chem. Mater., 2020, 32, 4600–4608 CrossRef CAS.
  56. H. Chaudhary, I. A. Iashchishyn, N. V. Romanova, M. A. Rambaran, G. Musteikyte, V. Smirnovas, M. Holmboe, C. A. Ohlin, Z. M. Svedruzic and L. A. Morozova-Roche, ACS Appl. Mater. Interfaces, 2021, 13, 26721–26734 CrossRef CAS PubMed.
  57. E. M. Villa, C. A. Ohlin, E. Balogh, T. M. Anderson, M. D. Nyman and W. H. Casey, Angew. Chem., Int. Ed., 2008, 47, 4844–4846 CrossRef CAS PubMed.
  58. E. M. Villa, C. A. Ohlin, E. Balogh, T. M. Anderson, M. D. Nyman and W. H. Casey, Am. J. Sci., 2008, 308, 942–953 CrossRef CAS.
  59. E. M. Villa, C. A. Ohlin and W. H. Casey, J. Am. Chem. Soc., 2010, 132, 5264–5272 CrossRef CAS PubMed.
  60. E. M. Villa, C. A. Ohlin, J. R. Rustad and W. H. Casey, J. Am. Chem. Soc., 2009, 131, 16488–16492 CrossRef CAS PubMed.
  61. J. M. Jehng and I. E. Wachs, Chem. Mater., 1991, 3, 100–107 CrossRef CAS.
  62. M. S. P. Francisco and Y. Gushikem, J. Mater. Chem., 2002, 12, 2552–2558 RSC.
  63. S. Morselli, P. Moggi, D. Cauzzi and G. Predieri, in Stud. Surf. Sci. Cat, ed. B. Delmon, P. A. Jacobs, R. Maggi, J. A. Martens, P. Grange and G. Poncelet, Elsevier, 1998, vol. 118, pp. 763–772 Search PubMed.
  64. C. C. Perry, X. Li and D. N. Waters, Spectrochim. Acta, Part A, 1991, 47, 1487–1494 CrossRef.

Footnote

Electronic supplementary information (ESI) available: Details related to synthesis and film deposition, Raman, PXRD, ellipsometry, 17O-NMR, EDS, SEM and AFM measurements. See DOI: 10.1039/d1dt03116c

This journal is © The Royal Society of Chemistry 2021
Click here to see how this site uses Cookies. View our privacy policy here.