Next Article in Journal
Impact of the Subunit Arrangement on the Nonlinear Absorption Properties of Organometallic Complexes with Ruthenium(II) σ-Acetylide and Benzothiadiazole as Building Units §
Previous Article in Journal
Transmetalation from Magnesium–NHCs—Convenient Synthesis of Chelating π-Acidic NHC Complexes
Previous Article in Special Issue
High-Pressure Routes to New Pyrochlores and Novel Magnetism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Titanium Doping of the Metallic One-Dimensional Antiferromagnet, Nb12O29

1
Department of Chemistry, University College London, Christopher Ingold Building, 20 Gordon Street, London WC1H 0AJ, UK
2
School of Physical Sciences, Ingram Building, University of Kent, Canterbury, Kent CT2 7NH, UK
*
Author to whom correspondence should be addressed.
Submission received: 4 March 2019 / Revised: 7 May 2019 / Accepted: 21 May 2019 / Published: 23 May 2019
(This article belongs to the Special Issue Magnetic Oxide Materials)

Abstract

:
Monoclinic Nb12O29 undergoes a charge ordering transition to form antiferromagnetic Nb4+ chains (TN ~ 12 K) spaced 15.7 Å apart, which are coupled through mediation from a subset of metallic electrons which are present over all temperature regimes. We present the effects of disrupting the delicate electronic equilibrium in monoclinic Nb12O29 through doping Nb4+ (d1) with Ti4+ (d0) ions in the series, TixNb12−xO29. Powder neutron diffraction demonstrates that Ti is distributed over all of the 6 crystallographically distinct Nb positions. Magnetic susceptibility measurements reveal a rapid suppression of the magnetic ordered state on Ti doping, with a 3% percolation threshold consistent with the existence of one-dimensional Nb4+ chains. The reduction of the number of unpaired electrons on Ti4+ doping is shown to depopulate both localised and itinerant electron subsets, demonstrating that they are intrinsic to the properties of the system, which is argued to be a direct consequence of the mixture of bonding schemes within the lattice.

1. Introduction

Charge ordering is a property of mixed valent solids, where below a temperature defined as TCO, different electronic configurations adopt distinct crystallographic arrangements. Charge ordering has become an important issue in understanding the electronic properties of transition metal oxides. Prominent examples include the ordering of Mn3+ and Mn4+ ions in perovskite manganates, creating a localised electronic state that is in competition with the ferromagnetic metallic state associated with colossal magnetoresistance. These competing interactions result in compositions that display a transition between the two states as a function of temperature [1]. The metal-insulator transition in Fe3O4 lead Verwey to predict the ordering of Fe2+ and Fe3+ ions within the spinel lattice [2], which was the first realisation of this phenomena and the extent to which it can influence the electronic properties of strongly correlated electronic systems. A related property found in many ceramics is that of cation ordering of different atoms within a lattice. For example, Ba3ZnTa2O9 contains ordered Zn and Ta ions in the perovskite framework [3], the extent of which critically affects the dielectric response in the material. It has been suggested that the gross differences in size and charge are the controlling factors behind the ordering of ions that may otherwise reside on crystallographically-equivalent sites. Charge ordering is the electronic analogue of cation ordering; it is a consequence of electron rather than atomic ordering, and can occur in extremely complex arrangements, which is demonstrated by the charge ordering of Ir3+/Ir4+ ions in CuIr2S4 [4] or the Mn3+/Mn4+ ions in LiMn2O4 [5].
Monoclinic Nb12O29 or Nb4+2Nb5+10O29 has novel metallic and low dimensional antiferromagnetic properties, despite its relatively low numbers of unpaired electrons [6,7,8,9], similar to its orthorhombic polymorph [10]. Its crystallographic shear plane structure is composed of (4 × 3) blocks of corner-shared NbO6 octahedra, which are edge shared to the next (4 × 3) block in a diagonal arrangement in the xz plane, shown in Figure 1.
The long-ranged, magnetically ordered state is accomplished through a charge ordering transition [8,11]. The localisation of Nb4+ ions (d1, S = 1/2) occurs on a specific crystallographic site, amidst neighbouring diamagnetic Nb5+ (d0, S = 1/2) ions, which repeats down the y-axis, giving rise to the formation of chains. The one-dimensional chains of localised Nb4+ (S = 1/2) order in three dimensions owing to interactions with a subset of delocalised electrons. Zero-field μSR and magnetic susceptibility measurements confirm the presence of a long-range antiferromagnetic state, rather than the more probable spin glass configuration for magnetic impurities embedded in a sea of conduction electrons [11]. Nb12O29 is an important example of how a low dimensional magnetic structure, with chains of exceptional distances apart (10.5, 15.1 and 15.7 Å), can be derived from a condensed three-dimensional structure. An interesting comparison can be drawn with the magnetic interactions in Nb12O29 and those of organic magnets. The reason for this is that for true low dimensional character, the magnetic exchange needs to be very anisotropic (e.g., for 1D systems, Jchain >>> Jinterchain), and therefore, to possess different exchange lengths. Organic magnets can provide suitable separation through the inclusion of bulky organic ligands, which is not possible in dense condensed matter systems. The low temperature structure of monoclinic Nb12O29, comprises of Nb4+ (S = 1/2) chains with distances of at least 10.5 Å apart and represent an important example of the unusual electronic properties that can be achieved when mixed d1 (S = 1/2) and diamagnetic d0 ions occupy the same lattice. The susceptibility of the occurrence of reside on different crystallographic sites is provided by the different environment that the two electronic configurations prefer. Nb4+ (d1) ions tend to undergo an out-of-centre distortion, providing one short, one long and 4 equal bonds in the octahedral environment, to which d1 ions are less susceptible; therefore, systems can often lower their energy at low temperature by arranging these different configurations onto distinct crystallographic sites. The particular relevance of this in terms of the electronic structure is that ordered systems containing mixed d0 and d1 ions are composed of closed and open shell electronic configurations, respectively, and offer the opportunity of producing unique electronic ground states.
Long ranged antiferromagnetic order is not commonly observed concomitantly with metallic properties, as the formation of antiparallel spins normally prohibits free movement of electrons, although there are many cases where there is a subtle balance between antiferromagnetic localised character and metallic itinerant properties. This is particularly pertinent in the electronic properties of high temperature superconducting cuprates, where the destruction of antiferromagnetic order with chemical doping can result in a transition to a superconducting state. Niobium oxides provide an interesting analogy to the high TC cuprates, both NbO [12] and AxNbO2 [13] undergo superconducting transitions. It might be expected that Nb12O29 show such a transition through similar strong metal–metal bonding, though the formation of localised magnetic electrons alongside metallic electrons may tend to suppress the superconducting state. Therefore, investigations of doped Nb12O29 are important in the study of the limit of the antiferromagnetic state, and to investigate whether suppression of the long ranged magnetic state results in superconductivity, as observed in certain cuprates. Furthermore, it is interesting to note the symmetric nature of the electronic configuration between these niobates and high TC cuprates, with Nb4+ possessing d1 (S = 1/2) and the high temperature superconducting cuprates based on doped d9 (S = 1/2) systems.
The origin of the magnetism in Nb12O29 has been extensively discussed. First principal calculations have suggested a Stoner instability attributed to Nb ions close to a valence of 5+ [14], whereas a more recently study has found the presence of low dimensional spin chains [15]. In this paper, we show that doping of very small amounts of Nb4+ (d1) with diamagnetic Ti4+ (d0) ions in Nb12O29 destroys the long-range magnetic order, revealing a low percolation threshold of around 3%, consistent with the presence of one-dimensional antiferromagnetic magnetic ordering. Both the metallic conduction and balance between localised and itinerant electrons remains with low concentrations of Ti, providing strong evidence that the electrons involved in the one-dimensional antiferromagnetic ordering and metallic properties originate from distinct electronic subsets.

2. Results and Discussion

Compositions with the formulae TiNb11O29 and Ti2Nb10O29 have both been previously reported [16,17,18], making the formation of a TixNb12−xO29 (0 ≤ x ≤ 2) solid solution likely. The TixNb12−xO29 series formed single phase materials with the monoclinic Nb12O29 structure throughout the range 0 < x < 1, although requiring different synthesis conditions. The volume of the resulting unit cell, shown in Figure 2a, was found to vary smoothly with composition, confirming homogeneous inclusion of the different sized Ti ions into the lattice. SQUID magnetometry measurements showed an equally consistent reduction in the Curie constant, C and Weiss Temperature, θ, as obtained from the fitting of the magnetic susceptibility to the Curie-Weiss relationship in the paramagnetic regime, further demonstrating the homogeneous inclusion of Ti as well as its direct effect on the amount of localised moment, as shown in Figure 2b,c.
Powder neutron diffraction was collected on one composition, TiNb11O29, at room temperature to give an estimation of the distribution of the Ti and Nb atoms within the lattice. Powder X-ray diffraction data was well fitted in the A 2/m space group, as determined for monoclinic Nb12O29 at room temperature [8], which was confirmed with neutron diffraction and shown in Figure 3. Rietveld refinement of the neutron powder diffraction pattern of the TixNb12−xO29 series is particularly accurate in determining the composition of the 6 crystallographic inequivalent metal sites, due to the contrasting coherent neutron scattering lengths of −3.438 fm (Ti) and 7.054 fm (Nb). A previous neutron study of Ti2Nb12O29 has shown the Ti atoms to show a slight preference to be located on the edges of the (4 × 3) block [18]. In contrast, the fractional occupancies in these refinements showed the Ti and Nb in TiNb11O29 to be evenly distributed over all site. The resulting atomic coordinates are provided in Table 1. The difference in the results from the two neutron patterns can be accounted for by the very different synthetic techniques used. The synthesis of Ti2Nb10O29, containing only d0 ions, was reported to be formed in normal atmospheric conditions over long synthesis periods, thereby achieving the thermodynamically most stable product [18].Through Rietveld refinement of a large spectrum of samples formed under different synthetic conditions, we have demonstrated that the most crystalline samples are produced through a rapid synthesis procedure at extremely hot temperature [7]. A typical route for Ti containing Nb12O29 would involve temperatures of 1280 °C for 75 min, followed by rapid quenching, which would tend to produce a kinetically stable product and does not allow the Ti and Nb to order in the most energetically favourable manner. From these arguments, and as all samples are produced under similar synthetic conditions, other compositions are assumed to have the same even distribution of Ti ions.
Figure 4 shows the magnetic susceptibility of three members of the TixNb12−xO29 series below 25 K. It demonstrates that the temperature of the magnetic transition, as determined from the point of steepest gradient below the maximum, is extremely sensitive to composition. The Neel temperature of approximately 12 K for Nb12O29 is shifted down to around 8.5 K for Ti0.05Nb11.95O29. An increase in the Ti concentration to Ti0.1Nb11.9O29 further reduces the ordering temperature to around 6.5 K. At low temperatures, Ti0.1Nb11.9O29 and compositions with Ti content greater than 0.1 show an additional increase in the susceptibility below the ordering temperature. The appearance of this Curie-tail is characteristic of doped low dimensional magnetic systems. The inclusion of the non-magnetic Ti4+ impurities, although amounting to less than 1% of the metal content in the lattice, fragments the alignment of Nb4+ spins and hinders the formation of a long range ordered state. Compositions with x > 0.3 show purely Curie-Weiss behaviour over the entire temperature range, and no transition to a long ranged ordered state is seen. A summary of the antiferromagnetic ordering temperatures across the series is given in Figure 5a, showing the consistent suppression of ordering temperature with Ti doping. These measurements are limited to temperature of 1.8 K and above; however, it is clear from extrapolation of TN across the series that the loss of long range magnetic order occurs at x ~ 0.35, corresponding to around 3% doping level.
The amount of non-magnetic dopant required to destroy long range magnetic ordering is crucially dependent on the dimensionality of the magnetic structure. Three-dimensional systems, where the exchange energy is approximately equal in all three spatial dimensions, are relatively resilient to non-magnetic impurities. Two-dimensional, long-range magnetic systems, where the exchange energy is greater within planes than in the third direction, is destroyed at lower concentrations than in the three-dimensional case. One-dimensional magnetic systems, where the exchange energy along one direction, forming chains, is greater than in the other two, can only tolerate very small changes in the stoichiometry. The destruction of magnetic order with dopant ions is related to the percolation threshold, which changes with dimensionality and structure. A two-dimensional square lattice has a percolation threshold of 0.5, whereas the theoretically percolation threshold for a one-dimensional magnet is 1, that is, there is no tolerance for non-magnetic impurities. There are a number of example of this in the literature, the one-dimensional spin-Peierls transition in CuGeO3 can be destroyed by doping of only 2% Zn on the copper site [19]. Similarly the transition in NaxV2O5 to a low dimensional zigzag arrangement [20] is destroyed when x ≤ 0.97, which effectively introduces 3% of non-magnetic V5+ into the V4+ chains [21]. The low percolation threshold for Ti doped Nb12O29 (around 3%), is consistent with the one-dimensional nature of the magnetic structure.
Figure 5b shows the temperature dependence of the resistivity of TixNb12−xO29 (x = 0.0, 0.05, 0.1 and 0.2). As these experiments were performed on powdered samples, the effect of grain boundaries can vary from one experiment to another, thereby undermining the reliable of the absolute values. However, all the samples show metallic behaviour and exhibit a general trend for the resistivity to reduce with the amount of Ti doping. This result may initially seem somewhat counterintuitive, as substitution of Nb4+ with Ti4+ reduces the number of unpaired electrons in the system and the presence of defects reduces the Nb–O and Nb–Nb overlap; however, the basis for this may lie in the interaction between the localised and itinerant electrons. The increase in the concentration of Ti4+ ions also reduces the number of localised electrons and, therefore, the scattering between the itinerant and localised electrons, whilst the depopulation of the conduction band and lowering of the Fermi energy may not necessarily have a significant effect at such low values of x.
The origin of the localised and itinerant electrons is an important issue in the electronic properties of these materials, which is governed by the complex bonding scheme in the monoclinic block structure. The retention of the balance between localised and delocalised electrons is in stark contrast to the properties of many charge-ordered systems such as Fe3O4 or the manganate perovskites. Here, the localised and delocalised properties are in direct competition and the charge ordering causes all of the electrons to localise that were previously itinerant, thereby causing a metal-insulator transition. In the case of TixNb12−xO29, the electrons can be thought of as distinct subsets of electrons that are intrinsic to the structure. A unit cell of the (Ti, Nb)–O structure is shown in Figure 6. It contains corner-sharing octahedra where the Nb–Nb bond lengths are around 3.7 Å and separated by an oxygen atom, with an Nb–O–Nb bond angle of around 180°. In contrast, the edge- shared octahedra contain Nb–O–Nb bond angles of approximately 90°, producing Nb–Nb bond lengths of around 3.18 Å. These two bonding schemes will clearly cause differing degrees of Nb–Nb and Nb–O overlap within the lattice. The octahedra in the centre of the (4 × 3) block contains only corner-sharing configurations, thereby leading to limited overlap and explaining the localised nature of the Nb4+ ions located at this site. In contrast, the relatively short Nb–Nb distances in the outer octahedra of the (4 × 3) block can give rise to direct Nb–Nb overlap, forming a conduction band and resulting in the presence of itinerant electrons. The inclusion of Ti ions in the lattice equally reduces the concentration of both the localised and delocalised electrons, as the underlying combination of edge and corner shared octahedra remains the same.

3. Materials and Methods

Compositions of TixNb12−xO29 (x = 0.0, 0.05, 0.1, 0.2, 0.25, 0.4, 0.5, 0.6, 0.8 and 1.0) were synthesised from stoichiometric amounts of Nb, H-Nb2O5 and TiO2. TiO2 was dried before use at 700 °C for 24 h. Commercially available Nb2O5 was heated at 1100 °C for 12 h before use to convert it to H-Nb2O5. The starting materials were thoroughly ground together, made into pellets, and sealed in an evacuated quartz tube. The heating temperatures and times were found to be critical in the synthesis of single-phase materials, which was further sensitive to composition. Samples were heated to high temperature, as described in Table 2, and then removed from the furnace and dropped into cold water to quench the reaction. The phase purity of various products under different synthetic conditions was determined by Rietveld refinement of powder X-ray diffraction, obtained from a Siemens D500 diffractometer (Siemens now Bruker, Karlsruhe, Germany). The optimal synthetic conditions for different Ti compositions are summarised in Table 2, and a typical X-ray pattern is shown in Figure 7, to demonstrate the high crystallinity achieved with this method. The resulting lattice parameters determined from Rietveld refinement of powder X-ray diffraction data is provided in Table 3. Magnetic susceptibility measurements were performed using a Quantum Design MPMS 7 SQUID magnetometer (Quantum Design, San Diego, CA, USA). Resistivity measurements were performed in the standard four probe technique. Powder neutron diffraction was performed on the D2B diffractometer built at the Institut Laue Langevin, France at a wavelength, λ = 1.594 Å.

4. Conclusions

The synthesis and characterization of a series of magnetic oxides with the general formula TixNb12−xO29 support the hypothesis that this is a low-dimensional magnet created by the localisation of Nb4+ onto specific sites that are absent of metal–metal bonding. These results will greatly help with the understanding of recent electrochemical studies [22].

Author Contributions

I.K., J.E.L.W. and M.A.G. contributed equally to the synthesis and characterization. M.A.G. designed the experiment and wrote the paper.

Funding

This research was funded by the Royal Society (WM120049).

Acknowledgments

We thank the Institute Laue Langevin, France for provision of neutron beam time. We thank Alan Hewat for assistance with the experiments. We thank Manisha Patel with assistance with synthesis and Pawel Zajdel for useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rao, C.N.R.; Arulraj, A.; Santosh, P.N.; Cheetham, A.K. Charge-Ordering in Manganates. Chem. Mater. 1998, 10, 2714–2722. [Google Scholar] [CrossRef]
  2. Verwey, E.J.W. Electronic Conduction of Magnetite (Fe3O4) and its Transition Point at Low Temperatures. Nature 1939, 144, 327–328. [Google Scholar] [CrossRef]
  3. Jacobson, A.J.; Collins, B.M.; Fender, B.E.F. A powder neutron and X-ray diffraction determination of the structure of Ba3Ta2ZnO9: An investigation of perovskite phases in the system Ba–Ta–Zn–O and the preparation of Ba2TaCdO5.5 and Ba2CeInO5.5. Acta Crystallogr. B 1976, 32, 1083–1087. [Google Scholar] [CrossRef]
  4. Radaelli, P.G.; Horibe, Y.; Gutmann, M.J.; Ishibashi, H.; Chen, C.H.; Ibberson, R.M.; Koyama, Y.; Hor, Y.S.; Kiryukhin, V.; Cheong, S.W. Formation of isomorphic Ir3+ and Ir4+ octamers and spin dimerization in the spinel CuIr2S4. Nature 2002, 416, 155–158. [Google Scholar] [CrossRef] [PubMed]
  5. Rodríguez-Carvajal, J.; Rousse, G.; Masquelier, C.; Hervieu, M. Electronic Crystallization in a Lithium Battery Material: Columnar Ordering of Electrons and Holes in the Spinel LiMn2O4. Phys. Rev. Lett. 1998, 81, 4660–4663. [Google Scholar] [CrossRef]
  6. Norin, R.; Hope, H.; Nevald, R.; Frank, V.; Brunvoll, J.; Bunnenberg, E.; Djerassi, C.; Records, R. The Crystal Structure of Nb12O29 (mon). Acta Chem. Scand. 1966, 20, 871–880. [Google Scholar] [CrossRef]
  7. Waldron, J.E.L.; Green, M.A.; Neumann, D.A. Solid State Symposium, Materials Research Society. In Solid State Symposium, Materials Research Society; Greedan, J.E., Ed.; MRS: Boston, MA, USA, 2000. [Google Scholar]
  8. Waldron, J.E.L.; Green, M.A.; Neumann, D.A. Charge and spin ordering in monoclinic Nb12O29. J. Am. Chem. Soc. 2001, 123, 5833–5834. [Google Scholar] [CrossRef]
  9. Waldron, J.E.L.; Green, M.A.; Neumann, D.A. Structure and electronic properties of monoclinic Nb12O29. J. Phys. Chem. Solids 2004, 65, 79–86. [Google Scholar] [CrossRef]
  10. Cava, R.J.; Batlogg, B.; Krajewski, J.J.; Gammel, P.; Poulsen, H.F.; Peck, W.F.; Rupp, L.W. Antiferromagnetism and Metallic Conductivity in Nb12O29. Nature 1991, 350, 598–600. [Google Scholar] [CrossRef]
  11. Lappas, A.; Waldron, J.E.L.; Green, M.A.; Prassides, K. Magnetic ordering in the charge-ordered Nb12O29. Phys. Rev. B 2002, 65, 134405. [Google Scholar] [CrossRef]
  12. Hulm, J.K.; Jones, C.K.; Hein, R.A.; Gibson, J.W. Superconductivity in the TiO and NbO Systems. J. Low Temp. Phys. 1972, 7, 291–307. [Google Scholar] [CrossRef]
  13. Geselbracht, M.J.; Richardson, T.J.; Stacy, A.M. Superconductivity in the Layered Compound LixNbO2. Nature 1990, 345, 324–326. [Google Scholar] [CrossRef]
  14. Fang, C.M.; Van Huis, M.A.; Xu, Q.; Cava, R.J.; Zandbergen, H.W. Unexpected origin of magnetism in monoclinic Nb12O29 from first-principles calculations. J. Mater. Chem. C 2015, 3, 651–657. [Google Scholar] [CrossRef]
  15. Lee, K.W.; Pickett, W.E. Organometalliclike localization of 4d-derived spins in an inorganic conducting niobium suboxide. Phys. Rev. B 2015, 91. [Google Scholar] [CrossRef]
  16. Wadsley, A.D. Mixed oxides of titanium and niobium. I. Acta Crystallogr. 1961, 14, 660–664. [Google Scholar] [CrossRef]
  17. Wadsley, A.D. Mixed oxides of titanium and niobium. II. The crystal structures of the dimorphic forms Ti2Nb10O29. Acta Crystallogr. 1961, 14, 664–670. [Google Scholar] [CrossRef]
  18. Dreele, R.B.V.; Cheetham, A.K. The Structures of Some Titanium–Niobium Oxides by Powder Neutron Diffraction. Proc. Royal Soc. A 1974, 338, 311–326. [Google Scholar] [CrossRef]
  19. Hase, M.; Koide, N.; Manabe, K.; Sasago, Y.; Uchinokura, K.; Sawa, A. Antiferromagnetic Long-Range Order caused by Nonmagnetic Impurities; Magnetization of Single-Crystal Cu1−xZnxGeO3. Physica B 1995, 215, 164–170. [Google Scholar] [CrossRef]
  20. Van Smaalen, S.; Daniels, P.; Palatinus, L.; Kremer, R.K. Orthorhombic versus monoclinic symmetry of the charge-ordered state of NaV2O5. Phys. Rev. B 2002, 65, 060101. [Google Scholar] [CrossRef]
  21. Isobe, M.; Ueda, Y. Effect of Na-deficiency on the spin-Peierls transition in α′-NaV2O5. J. Alloys Compd. 1997, 262–263, 180–184. [Google Scholar] [CrossRef]
  22. Wu, X.; Miao, J.; Han, W.; Hu, Y.-S.; Chen, D.; Lee, J.-S.; Kim, J.; Chen, L. Investigation on Ti2Nb10O29 Anode Material for Lithium-ion Batteries. Electrochem. Commun. 2012, 25, 39. [Google Scholar] [CrossRef]
Figure 1. Crystallographic shear structure of monoclinic Nb12O29, composed of (4 × 3) blocks of NbO6 octahedral units. Green polyhedral represent Nb5+ (d0) octahedral units, where the Nb4+ (d1) octahedral units forming chains down the y axis are shown in blue.
Figure 1. Crystallographic shear structure of monoclinic Nb12O29, composed of (4 × 3) blocks of NbO6 octahedral units. Green polyhedral represent Nb5+ (d0) octahedral units, where the Nb4+ (d1) octahedral units forming chains down the y axis are shown in blue.
Inorganics 07 00066 g001
Figure 2. The variation of the (a) volume of unit cell (b) Curie constant, C and (c) Weiss temperature, θ, as a function of Ti doping in TixNb12−xO29, demonstrating the homogeneous inclusion of the Ti ions into the Nb12O29 lattice.
Figure 2. The variation of the (a) volume of unit cell (b) Curie constant, C and (c) Weiss temperature, θ, as a function of Ti doping in TixNb12−xO29, demonstrating the homogeneous inclusion of the Ti ions into the Nb12O29 lattice.
Inorganics 07 00066 g002
Figure 3. Rietveld refinement of powder neutron diffraction data of TiNb11O29 at room temperature.
Figure 3. Rietveld refinement of powder neutron diffraction data of TiNb11O29 at room temperature.
Inorganics 07 00066 g003
Figure 4. Magnetic Susceptibility from 2–25 K of Ti0.05Nb11.95O29, Ti0.1Nb11.9O29 and TiNb11O29, showing a rapid suppression of the antiferromagnetic ordering, as well as an increase in the Curie tail, on increasing Ti concentration.
Figure 4. Magnetic Susceptibility from 2–25 K of Ti0.05Nb11.95O29, Ti0.1Nb11.9O29 and TiNb11O29, showing a rapid suppression of the antiferromagnetic ordering, as well as an increase in the Curie tail, on increasing Ti concentration.
Inorganics 07 00066 g004
Figure 5. The (a) antiferromagnetic ordering temperatures and (b) resistivity of selected members of the TixNb12−xO29 series showing the rapid suppression of long ranged magnetic behaviour with Ti doping.
Figure 5. The (a) antiferromagnetic ordering temperatures and (b) resistivity of selected members of the TixNb12−xO29 series showing the rapid suppression of long ranged magnetic behaviour with Ti doping.
Inorganics 07 00066 g005
Figure 6. Unit cell of monoclinic, Nb12O29, showing contrasting bonding of niobium sites on the edge of the (4 × 3) block, which possess strong metal–metal bonding at distances of ~3.2 Å, whereas those in the middle are absent of direct Nb–Nb interactions.
Figure 6. Unit cell of monoclinic, Nb12O29, showing contrasting bonding of niobium sites on the edge of the (4 × 3) block, which possess strong metal–metal bonding at distances of ~3.2 Å, whereas those in the middle are absent of direct Nb–Nb interactions.
Inorganics 07 00066 g006
Figure 7. X-ray powder diffraction of TiNb11O29, demonstrating the highly crystalline oxides formed through a rapid high temperature reaction.
Figure 7. X-ray powder diffraction of TiNb11O29, demonstrating the highly crystalline oxides formed through a rapid high temperature reaction.
Inorganics 07 00066 g007
Table 1. Atomic position and site occupancy of TiNb11O29 at room temperature, as determined from Rietveld refinement of powder neuron diffraction data in the monoclinic A 2/m space group.
Table 1. Atomic position and site occupancy of TiNb11O29 at room temperature, as determined from Rietveld refinement of powder neuron diffraction data in the monoclinic A 2/m space group.
xyz
Nb/Ti (1)0.0973 (15)00.0760 (14)
Nb/Ti (2)0.0906 (18)00.6839 (17)
Nb/Ti (3)0.1098 (13)00.8781 (12)
Nb/Ti (4)0.3749 (12)00.1507 (10)
Nb/Ti (5)0.3598 (14)00.7633 (15)
Nb/Ti (6)0.3502 (17)00.9554 (15)
O (13)1/200
O (14)0.0660 (17)00.1740 (16)
O (15)0.0659 (19)00.3545 (14)
O (16)0.0752 (15)00.5872 (14)
O (17)0.0669 (14)00.7711 (14)
O (18)0.0842 (14)00.9735 (12)
O (19)0.2187 (14)00.0974 (14)
O (20)0.2315 (16)00.7312 (15)
O (21)0.2155 (16)00.9206 (14)
O (22)0.3706 (14)00.0579 (14)
O (23)0.3559 (17)00.2515 (15)
O (24)0.3554 (17)00.4445 (13)
O (25)0.3609 (16)00.6696 (17)
O (26)0.3619 (15)00.8555 (14)
O (27)0.4919 (16)00.1863 (14)
Table 2. Heating conditions for optimal synthesis of the TixNb12−xO29 series.
Table 2. Heating conditions for optimal synthesis of the TixNb12−xO29 series.
ProductTemperature/°CHeating Time/min
Nb12O29128075
Ti0.05Nb11.95O29125075
Ti0.1Nb11.9O29125075
Ti0.2Nb11.8O29125060
Ti0.25Nb11.75O29125060
Ti0.4Nb11.6O29125050
Ti0.5Nb11.5O29125050
Ti0.6Nb11.4O29125045
Ti0.8Nb11.2O29125045
TiNb11O29125040
Table 3. Unit cell parameters of the TixNb12−xO29 series as determined from X-ray powder diffraction.
Table 3. Unit cell parameters of the TixNb12−xO29 series as determined from X-ray powder diffraction.
Samplea (Å)b (Å)c (Å)β (°)Volume (Å3)
Nb12O2915.6920 (7)3.8303 (2)20.7171 (9)113.112 (3)1145.26 (2)
Ti0.05Nb11.95O2915.685 (4)3.8306 (8)20.706 (4)113.10 (1)1144.3 (6)
Ti0.1Nb11.9O2915.682 (3)3.8297 (3)20.702 (3)113.09 (2)1143.7 (4)
Ti0.2Nb11.8O2915.673 (3)3.8278 (7)20.703 (4)113.08 (1)1142.7 (5)
Ti0.25Nb11.75O2915.673 (3)3.8286 (7)20.692 (4)113.10 (1)1142.1 (5)
Ti0.4Nb11.6O2915.6516 (3)3.8279 (1)20.6699 (4)113.088 (14)1139.18 (4)
Ti0.5Nb11.5O2915.6431 (2)3.8277 (1)20.6606 (2)113.059 (7)1138.19 (2)
Ti0.6Nb11.4O2915.6351 (2)3.8280 (1)20.6529 (2)113.056 (8)1137.3 (2)
Ti0.8Nb11.2O2915.6004 (2)3.8253 (4)20.6191 (3)113.039 (9)1132.29 (2)
TiNb11O2915.593 (3)3.8221 (6)20.612 (4)113.01 (1)1130.7 (5)

Share and Cite

MDPI and ACS Style

Kruk, I.; Waldron, J.E.L.; Green, M.A. Titanium Doping of the Metallic One-Dimensional Antiferromagnet, Nb12O29. Inorganics 2019, 7, 66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050066

AMA Style

Kruk I, Waldron JEL, Green MA. Titanium Doping of the Metallic One-Dimensional Antiferromagnet, Nb12O29. Inorganics. 2019; 7(5):66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050066

Chicago/Turabian Style

Kruk, Izabella, Joanna E. L. Waldron, and Mark A. Green. 2019. "Titanium Doping of the Metallic One-Dimensional Antiferromagnet, Nb12O29" Inorganics 7, no. 5: 66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050066

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop