Elsevier

Acta Materialia

Volume 206, March 2021, 116636
Acta Materialia

Lattice evolution, ordering transformation and microwave dielectric properties of rock-salt Li3+xMg2–2xNb1-xTi2xO6 solid-solution system: A newly developed pseudo ternary phase diagram

https://doi.org/10.1016/j.actamat.2021.116636Get rights and content

Abstract

New types of multi-component Li3+xMg2–2xNb1-xTi2xO6 (0 ≤ x ≤ 1) solid-solution ceramics were designed based on the Li2TiO3−Li3NbO4−MgO pseudo ternary phase diagram and studied for microwave dielectric applications. As the substitution amount (x) increased, we detected the phase transitions among the orthorhombic, cubic, and monoclinic phase driven by the compositional changes, as well as accompanied by an order-disorder-order transformation. A full range of solid solutions was formed between the Li3Mg2NbO6 and Li2TiO3 endmembers, with no trace of other impurities. In the sample with the low substitution concentration (x = 0.2 mol), a coherent phase interface (CPI) between the cubic and orthorhombic lattices was formed with no obvious misfit dislocation or stacking fault, indicating the small differences in crystal configuration, chemical bonding properties, and subcell lattice parameters between the two phases. Besides, there were observed diverse reconstructed superlattices, one kind is that possessed a “transition form” of the two phases and was formed nearby the CPI, and the other kind was formed based on the cubic or orthorhombic lattices independently and was observed on a larger scale nearby the CPI. The preferential substitutions of the non-equivalent cations, which were determined by ionic radius, electronegativity, and local electroneutrality, and the interfacial strains would together act on the formation of these superlattices. The Q × f values measured in the microwave range increased considerably around the compositional range where the superlattices were formed, indicating that the effect of reconstructed superlattices on the intrinsic loss should not be overlooked. As proven by the dielectric response in the high-frequency range (0.5 − 1 THz), the x = 0.2 sample indeed showed extremely higher Q × f values than other ones, which illustrated that the sample with the reconstructed superlattices was related to a small lattice vibrational anharmonicity that is favorable for the low dielectric loss.

Introduction

The ever-growing traffic explosion in mobile communications has recently drawn increasing interest in the designs of new microwave dielectric components for high data rates. Certain frequency range near the bottom of the millimeter-wave spectrum (30−80 GHz) is being used in the 5th generation cellular networks because of the enhanced spectrum bandwidth, which enables high-speed signal transmission [1]. Because the relative permittivity is inversely proportional to the signal propagation velocity through the medium, materials with low permittivities (εr≤25) are considered fitting for the millimeter-wave applications [2]. Besides, a high quality factor (Q) and a near-zero temperature coefficient of the resonant frequency (τf) are essential factors for the practical usage of dielectric ceramics [3,4].

In recent years, numerous ceramics with the rock-salt structure were explored with properties well-suited for millimeter-wave applications because of their low relative permittivities and low dielectric losses, such as Li2TiO3, Li3NbO4, Li3Mg2NbO6, and Li2Mg3TiO6, etc. [5], [6], [7], [8]. Among these matrix ceramics, the Li2TiO3 ceramic attracted much attention because of its unique positive τff=30−36 ppm/ °C) within the group of low-permittivity dielectric ceramics. According to the mixing rule of dielectrics, Li2TiO3 was a useful dopant to composite with another phase with a negative τf to get a near-zero τf value [9], [10], [11]. The dielectric loss will be inevitably increased if the impurity phase is introduced into the matrix, but this effect can be avoided by forming a solid solution in the Li2TiO3-based ceramic system, where a high Q together with a near-zero τf value can be jointly achieved. For instance, Bian et al. utilized Mg2+ ions to co-substitute Li+/Ti4+ions to form the (1-x)Li2TiO3-xMgO solid solutions, where the replacement mechanism is 3Mg2+⇄2Li++Ti4+, and the excellent microwave dielectric properties were achieved in the x = 0.24 sample with εr=19.2, Q × f = 106,226 GHz, and τf=3.56 ppm/ °C [12]. Interestingly, a continuous monoclinic-cubic phase transition accompanied by an order-disorder transformation occurred in the Li2TiO3-rich end of the Li2TiO3−MgO solid-solution system, and the transformed phase corresponded to the symmetry of the MgO cubic structure with Fm3¯m space group (S.G.). The Li2Mg3TiO6 ceramic adopted a disordered cubic structure and exhibited ultra-high Q × f values (εr=15.2, Q × f = 152,000 GHz, and τf=−39 ppm/ °C), and it belongs to the Li2TiO3−MgO solid-solution system [8]. Besides, phase-transition phenomena appeared to be common in the solid solutions formed by the endmembers with a rock-salt crystal configuration but different symmetries. For other examples of the two-endmember systems, an ordered cubic (S.G. I-43 m)-disordered cubic (S.G. Fm3¯m)-monoclinic (S.G. C2/c) phase transition was reported in the Li2TiO3−Li3NbO4 system, and an orthorhombic (S.G. Fddd)-cubic (S.G. Fm3¯m) phase transition was found in the Li3NbO4−MgO pseudo-binary system [13,14]. Among these solid solution systems, a wide range of desired properties was tunable by compositional modifications. On the whole, the above-mentioned studies indicated that the Li2TiO3, Li3NbO4, and MgO could form solid solutions in pairs. We supposed that the three-endmember solid solution systems would offer more degrees of freedom to obtain promising properties that cannot be achieved in the two-endmember systems. As we have preliminarily studied in our earlier work, the compound of the central section of the Li2TiO3−Li3NbO4−MgO phase diagram, Li5MgTiNbO8, adopted a pure cubic phase and exhibited excellent properties of εr=17.55, Q × f = 109,700 GHz, and τf=−32.5 ppm/ °C, which has shown great potential for the development of the materials with excellent performance in the Li2TiO3−Li3NbO4−MgO pseudo ternary system [15]. To make it easier to be comprehended the structural evolutions among the Li2TiO3, Li3NbO4, and MgO ceramics, we plotted the pseudo ternary phase diagram of the Li2TiO3−Li3NbO4−MgO system according to the data from the abovementioned literature and the other studies concerning the Li2MgTiO4, Li3Mg2NbO6, and Li4Mg3Ti2O9 compounds [7,16,17], as shown in Fig. 1.

Apart from the solid solutions in the Li2TiO3−Li3NbO4−MgO pseudo ternary system, there were many other two-endmember systems with large solid solubilities and excellent microwave dielectric properties, such as the Li2SnO3−MgO [18], Li2ZrO3−MgO [19], and Li2ZrO3−Li3NbO4 [20] systems. The universal large solid solubilities in the rock-salt solid solutions may be attributed to the loose constraint of the ion radius for the rock salt structure, where the radii of the cations (RA) and the radii of the anions (RX) should meet the scope of 0.42≤RA/RX≤0.72 [21]. A large solid solubility of a rock-salt system represents that there is a broad compositional range that can be adjusted to seek desired properties without worsening them due to impurities. On the other hand, it appeared that the composition-driven phase transitions in the abovementioned solid solutions were induced by the complex substitutions of the non-equivalent ions, and what could be determined was that the change of cation ordering and crystal symmetry would inevitably affect the intrinsic loss of the dielectrics [22,23]. But the studies concerning the mechanisms of the phase transitions and the origins of the ultra-low loss of these rock-salt systems were limited until now. It is noticeable that there has been a phenomenological correlation between the high ordering degree and high Q value widely studied in complex perovskites [24,25]. Besides, the high Q values were also found in the complex perovskites that were substituted by the non-equivalent ions, and the extremely low losses were owed to the formation of the ordering-induced domains in the samples with the low dopant or substituent concentration [26]. It is vital and necessary to understand the intrinsic loss affected by structural changes and the most promising approach is to study the high-frequency response of materials, including the whole submillimeter (0.3 − 3 THz) and part of the far-infrared (1.5 − 36 THz) range. The intrinsic dielectric properties of materials are overwhelmingly stronger than the extrinsic ones in the dielectric response at tremendously high frequency due to the proximity of phonon eigenfrequencies (generally in the order of 1011−1012 Hz) [27,28]. The classic damped oscillator model which fits the far-infrared reflectivity data were widely used to extrapolate the dielectric properties from far-infrared down to the microwave range, but it is not accurate for materials with permittivities below 20 [28]. Direct extrapolation of dielectric properties from submillimetre to microwave range is valid because the proportionality ε’’ (imaginary part of dielectric function) ∝ f (frequency) is roughly obeyed in the whole range from the microwave to submillimeter wave [27], [28], [29].

In this paper, we are aiming to clarify the relationship among the structural changes, phonon vibrations, and microwave dielectric properties of the Li3+xMg2–2xNb1-xTi2xO6 (0 ≤ x ≤ 1) ceramics in the Li2TiO3−Li3NbO4−MgO pseudo ternary system. The Li3Mg2NbO6 endmember was chosen due to its promising properties (εr=16.8, Q × f = 79,600 GHz and τf=−27.2 ppm/ °C), and several studies have proved that the properties could be effectively improved by the non-equivalent ion substitutions [7,[30], [31], [32]]. The Li2TiO3 endmember was selected because the temperature-stable sample was expected in the Li2TiO3-rich end of the phase diagram.

Section snippets

Material preparation

Li3+xMg2–2xNb1-xTi2xO6 (0 ≤ x ≤ 1) ceramics were prepared via a high-temperature solid-state reaction method. Reagent-grade raw powders of Li2CO3 (99.99%), Nb2O5 (99.99%), TiO2 (99.99%), and Mg(OH)2•4MgCO3•5H2O (99.95%) were baked at 150 #x00B0;C overnight to remove the moisture. The raw powders were weighed according to their stoichiometric ratio and ball milled with zirconia media in ethyl alcohol for 8 h and then were dried for 12 h. The dried slurries were calcined at 900 #x00B0;C − 1000 °C

Structural evolution and ordering transformation

The XRD patterns of the representative specimens of Li3+xMg2–2xNb1-xTi2xO6 (0 ≤ x ≤ 1) ceramics are shown in Fig. 3. Specifically, the patterns of x = 0 and x = 0.08 correspond to the pure orthorhombic phase (Li3Mg2NbO6: JCPDS no.36–1018) in the Fddd space group (No. 70). Yet an extra diffraction peak emerges on the left side of (026) peak in the pattern of x = 0.1, and the relative intensity of this peak increases with the increase of the x value in the range of 0.1 ≤ x ≤ 3.0. The specimens in

Conclusions

The Li3+xMg2–2xNb1-xTi2xO6 (0 ≤ x ≤ 1) solid-solution ceramics were designed by means of the Li2TiO3–Li3NbO4−MgO pseudo ternary phase diagrams and were synthesized via the standard solid-state reaction method. XRD patterns confirmed an infinite solid solution was formed in the full range of 0 ≤ x ≤ 1, where occurred the orthorhombic (S.G. Fddd)-cubic (S.G. Fm3¯m)-monoclinic (S.G. C2/c) composition-driven phase transition, as well as accompanied by an order-disorder-order transformation. Because

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant No. 51672038).

References (78)

  • J.X. Bi et al.

    Phase composition, microstructure and microwave dielectric properties of rock salt structured Li2ZrO3-MgO ceramics

    J Eur Ceram Soc

    (2018)
  • A. Salinas-Sanchez et al.

    Structural characterization of R2BaCuO5 (R= Y, Lu, Yb, Tm, Er, Ho, Dy, Gd, Eu and Sm) oxides by X-ray and neutron diffraction

    J Solid State Chem

    (1992)
  • A.E. Ringwood

    The Principles governing trace element distribution during magmatic crystallization Part I: The influence of Electronegativity

    Geochimica et Cosmochimica Acta

    (1955)
  • R.K. Singh

    Many body interactions in binary ionic solids

    Phys Rep

    (1982)
  • J.J. Bian et al.

    Structural evolution, grain growth kinetics and microwave dielectric properties of Li2Ti1-x(Mg1/3Nb2/3)xO3

    J Eur Ceram Soc

    (2018)
  • G. Wang et al.

    Synthesis, crystal structure and low loss of Li3Mg2NbO6 ceramics by reaction sintering process

    Ceram Int

    (2019)
  • T. Nakazawa et al.

    High energy heavy ion induced structural disorder in Li2TiO3

    Journal of Nuclear Materials

    (2007)
  • N.-.X. Xu et al.

    Structural evolution and microwave dielectric properties of MgO–LiF co-doped Li2TiO3 ceramics for LTCC applications

    Ceram Int

    (2014)
  • S.-.Y. Choi et al.

    Sintering kinetics by structural transition at grain boundaries in barium titanate

    Acta Mater

    (2004)
  • M. Amsif et al.

    Influence of rare-earth doping on the microstructure and conductivity of BaCe0.9Ln0.1O3−δ proton conductors

    J Power Sources

    (2011)
  • R. Amin et al.

    Part-II: exchange current density and ionic diffusivity studies on the ordered and disordered spinel LiNi0.5Mn1.5O4 cathode

    J Power Sources

    (2017)
  • L. Zhang et al.

    Influence of BaSnO3 additive on the energy storage properties of Na0. 5Bi0. 5TiO3-based relaxor ferroelectrics

    J Eur Ceram Soc

    (2018)
  • T.E. Hsieh et al.

    Experimental study of grain boundary melting in aluminum

    Acta Metallurgica

    (1989)
  • S.H. Yoon et al.

    Investigation of the relations between structure and microwave dielectric properties of divalent metal tungstate compounds

    J Eur Ceram Soc

    (2006)
  • H. Ohsato

    Origins of high Q on microwave tungstenbronze-type like Ba6−3xR8+2xTi18O54 (R: rare earth) dielectrics based on the atomic arrangements

    J Eur Ceram Soc

    (2007)
  • E. Li et al.

    Effects of Li2O-B2O3-SiO2 glass on the low-temperature sintering of Zn0.15Nb0.3Ti0.55O2 ceramics

    Ceram Int

    (2018)
  • M.T. Sebastian et al.

    Microwave Materials and Applications

    (2017)
  • H.-.H. Guo et al.

    Temperature stable Li2Ti0.75(Mg1/3Nb2/3)0.25O3-based microwave dielectric ceramics with low sintering temperature and ultra-low dielectric loss for dielectric resonator antenna applications

    J. Mater. Chem. C

    (2020)
  • J. Liang et al.

    Microwave Dielectric Properties of Li2TiO3 Ceramics Doped with ZnO–B2O3 Frit

    J. Am. Ceram. Soc.

    (2009)
  • D. Zhou et al.

    Microwave Dielectric Characterization of a Li3NbO4 Ceramic and Its Chemical Compatibility with Silver

    J. Am. Ceram. Soc.

    (2008)
  • L.L. Yuan et al.

    Microwave Dielectric Properties of the Lithium Containing Compounds with Rock Salt Structure

    Ferroelectrics

    (2009)
  • J. Bian et al.

    Structural Evolution and Microwave Dielectric Properties of Li(3−3x)M4xNb(1−x)O4 (M=Mg,Zn; 0≤x≤0.9)

    J. Am. Ceramic Society

    (2011)
  • L. Amaral et al.

    Grain growth anomaly and dielectric response in Ti-rich strontium titanate ceramics

    J. Phys. Chem. C

    (2013)
  • Z. Fang et al.

    Structure and microwave dielectric properties of the Li2/3(1−x)Sn1/3(1−x)MgxO systems (x = 0-4/7)

    J. Am. Ceram. Soc.

    (2018)
  • H. Yang et al.

    A new low-firing and high-Q microwave dielectric ceramic Li9Zr3NbO13

    J. Am. Ceram. Soc.

    (2018)
  • G.C. Mather et al.

    A review of cation-ordered rock salt superstructure oxides

    J Mater Chem

    (2000)
  • V.L. Gurevich et al.

    Intrinsic dielectric loss in crystals

    Adv Phys

    (1991)
  • E. Schlömann

    Dielectric Losses in Ionic Crystals with Disordered Charge Distributions

    Phys. Rev.

    (1964)
  • C.-.H. Wang et al.

    XRD and Raman Studies on the Ordering/Disordering of Ba(Mg1/3Ta2/3)O3

    J. Am. Ceram. Soc.

    (2009)
  • Cited by (53)

    View all citing articles on Scopus
    View full text