You are on page 1of 31

Journal Pre-proof

Investigation of Antimony ions doping on crystal structure and enhanced microwave


dielectric performance of MgTa2O6 ceramics

Liang Shi, Kui Liu, Cheng Liu, Dainan Zhang, Huaiwu Zhang

PII: S2352-8478(23)00027-8
DOI: https://doi.org/10.1016/j.jmat.2023.01.010
Reference: JMAT 667

To appear in: Journal of Materiomics

Received Date: 13 October 2022


Revised Date: 14 December 2022
Accepted Date: 25 January 2023

Please cite this article as: Shi L, Liu K, Liu C, Zhang D, Zhang H, Investigation of Antimony ions doping
on crystal structure and enhanced microwave dielectric performance of MgTa2O6 ceramics, Journal of
Materiomics (2023), doi: https://doi.org/10.1016/j.jmat.2023.01.010.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2023 Published by Elsevier B.V. on behalf of The Chinese Ceramic Society.


Jo
ur
na
lP
re
-p
ro
of
Investigation of Antimony ions doping on crystal structure and

enhanced microwave dielectric performance of MgTa2O6 ceramics


Liang Shi1, Kui Liu1, Cheng Liu1,2, Dainan Zhang1,2,
Huaiwu Zhang1,2*.
1
School of Electronic Science and Engineering, University of Electronic Science and
Technology of China, Chengdu, 610054, China.
2
State Key Laboratory of Electronic Thin Films and Integrated Devices, University of
Electronic Science and Technology of China, Chengdu, 610054, China.

f
r oo
Abstract -p
Dielectric ceramics are promising in large-scale commercial millimeter-wave
re
lP

communication technology, such as 5G and the upcoming 6G, thanks to their excellent

frequency selection characteristics and environmental stability. In this work, various


na

contents of Sb-O bonds were introduced into the MgTa2O6 lattice using the solid-phase
ur

reaction method to investigate the effects on the lattice and microwave dielectric
Jo

properties. XRD confirms that Sb ions successfully occupy Ta sites in the lattice and

cause lattice shrinkage and crystallinity deterioration, which leads to a slight decrease in

the quality factor. Furthermore, DFT calculations reveal that the doping leads to

electron-biased aggregation toward O atoms, causing higher Sb-O ionicity, but also

attenuates the degree of ionization of Ta and Mg ions, which makes the dielectric

constant of the doped samples vary non-monotonically with gradient doping.

Satisfactorily, Sb doping substantially enhanced the thermal stability of the ceramics,

with TCF values reduced from 36×10-6 oC-1 to 15×10-6 oC-1.

1
Keywords: MgTa2O6 ceramics; Dielectric; Raman; DFT calculation.

*Corresponding Author: Huaiwu Zhang, E-mail: hwzhang@uestc.edu.cn; Liang Shi:

shiliang_uestc@std.uestc.edu.cn.

1. Introduction

f
The extensive commercialization of new-generation 5G millimeter-wave

oo
communications and the innovative development of pre-research 6G communication

r
-p
technologies have propelled the development of millimeter wave circuits and devices to
re
a new level. As the carrier of millimeter wave circuits and devices, microwave dielectric
lP

ceramics (MWDCs) should have excellent dielectric properties and thermal stability,
na

which are the keys to meet the high frequency and integration of millimeter wave
ur

communication technology [1-4]. The heat loss generated in the transmission of


Jo

microwaves in circuits and devices, caused by dielectric loss, is more prominent than in

previous generations of communication technologies, aggravating power loss and signal

inaccuracy[5]; thus, low dielectric losses and high temperature tolerance are the current

research focus and impetus for microwave dielectric ceramics.[6-9]. Ta-based ceramics

are typically medium dielectric constant ceramics with favorable thermal conductivity

and dielectric properties, making them excellent prospects for use in high frequency

millimeter-wave communication devices [10, 11]. Sebastian et al. synthesized A5B4O15

(A=Ba, Sr, Mg, Ca, Zn; B=Nb, Ta) series ceramics with dielectric constants that varied

2
from 10 to 50, but the majority of them suffered unfavorable dielectric losses (Q×f < 40

000 GHz). The Mg- and Zn-rich AB2O6 (A=Ca, Mg, Mn, Co, Ni, Zn, and B=Nb, Ta)

series of dielectric ceramics demonstrate excellent microwave dielectric properties (εr >

20, Q×f > 60 000 GHz) and are excellent candidates for next-generation communication

technologies [10]. Among them, trirutile-type MgTa2O6 ceramics with appropriate

dielectric constants (εr ~ 28) and low dielectric losses (Q×f ~ 60 000 GHz) have drawn

f
extensive research. However, the ultra-high sintering temperature (over 1 500 °C) and

oo
poor temperature stability (~50×10-6 oC-1) of MgTa2O6 ceramics restrict their application

r
-p
in microwave circuits and devices [11-13]. Ion doping engineering is a widespread and
re
effective method being adopted for improving the properties of materials, for both
lP

organic and inorganic compounds [14, 15]. Several studies have found that doping with
na

low melting point oxides CuO [15], introducing non-stoichiometric ratios of MgO [10],
ur

and sintering aid Al2O3 [16] are effective in lowering the sintering temperature while
Jo

retaining tantalite's excellent microwave dielectric properties. However, the introduction

of such oxides causes the generation of heterogeneous phases, augments the complexity

of the phase control in the synthesis of the compounds, and deteriorates the microwave

dielectric properties. Consequently, a number of equivalent ion substitutions have been

spawned to preserve the phase constancy and optimize the microwave dielectric

properties of tantalite compounds. For example, Mn2+,Ni2+, Zn2+, and Co2+ were used

to partially substitute the occupant ions at the A-site in ZnTa2O6 and MgTa2O6

compounds to achieve an upgrade in quality factors or thermal stability[16-20], but few

3
of them achieved both thermal stability and an upgrade in quality factors. Due to the

limitations of the effect posited by A-site ion doping, researchers turned to explore the

effect of B-site ion doping. Nb ions, which belong to the same family as Ta elements,

were first introduced as B-site doping ions to achieve improvements in microwave

dielectric properties and thermal stability, but only trace amounts of Nb5+ were found to

cause the generation of the heterogeneous phase MgNb2O6[12, 21]. Intriguingly, a

f
positive promotion on quality factors was found after covalent substitution of Nb and Ta

oo
ions by Sb ions, for example, the introduction of 8% Sb at the Nb site in MgZrNb2O8

r
-p
ceramics reduced the dielectric loss[22], and the incorporation of 8% Sb in
re
Ag(Nb0.8Ta0.2)O3 ceramics effectively enhanced the thermal stability[23].
lP
na

Taking into account the merits of Sb ion doping and the lesser bond dissociation
ur

enthalpy (BDEs) of Sb-O bonds (434 kJ/mol) compared to Ta-O bonds (839 kJ/mol)[24].
Jo

We hypothesize that the introduction of partial Sb in MgTa2O6 ceramic compounds will

reduce the energy required to synthesize the trirutile phase, aiming to lower its initial

sintering temperature (1 550 °C), and enhance its pristine microwave dielectric

properties. In this work, Sb ions with 2%–12% variation are doped into MgTa2O6

ceramics using the solid-phase reactive sintering method. In combination with Raman

spectroscopy and first principle calculations, the effect of Sb ion doping on the crystal

structure, bond properties, and performance is revealed.

4
2. Results and discussion

2.1 lattice structure

The pure trirutile phase was confirmed in all doped samples by XRD diffraction

patterns, as shown in Figure 1. Remarkably, the positions and intensities of the

diffraction peaks of all the doped samples in the XRD patterns matched well in

f
comparison with the standard diffraction spectra (JCPDS 84-1679), and no peak

oo
splitting or heterogeneous peaks could be observed. Thus, observation of a single rutile

r
-p
phase demonstrates that the Sb ions enter the lattice smoothly and occupy the 4e
re
Wyckoff position within the doping limits in this work. In order to obtain more delicate
lP

changes in the crystal structure, we carried out structure refinement (GSAS software)
na

based on XRD data and acquired crystal structure data (cell parameters, chemical bond
ur

lengths) of different doped MgTa2O6 ceramics as shown in Table 1, Figure 2 and Figure
Jo

S1. The lattice parameters and bond lengths of Sb-doped MgTa2O6 ceramics are listed in

Table 1. As shown in Figure 2b, the lattice parameters (a, b, and V) gradually decrease

with increasing doping, which is primarily due to the smaller radius of the Sb5+ (0.6 Å)

compared to the Ta5+ (0.64 Å). However, the lattice parameter c, with a general tendency

to drop, shows a noticeable fluctuation, which is caused by the intermixing of the Ta-O

bond and Mg-O in the c direction, as shown in Figure 2c.

5
Table 1 Crystallographic parameters of the Sb-doped MgTa2O6 ceramics based on the XRD
refinement. (d is bond length, V is cell volume, Rwp and Rp is the refining reliability factor)

Doping a=b(Å) c (Å) V (Å3) d(Ta-O1 d(Ta-O2 d(Ta-O2 d(Mg-O d(Mg-O Rwp(%) Rp(%)

)×4 )×4 )’×4 1)×2 2)×4

x=0.02 4.717 6 9.212 5 205.035 2.021 2 1.978 8 2.031 0 2.010 1 2.049 6 6.01 3.65

x=0.04 4.717 2 9.214 4 205.04 2.026 4 1.981 9 2.026 7 2.027 1 2.059 1 5.31 3.73

f
oo
x=0.06 4.716 0 9.208 6 204.807 2.024 7 1.972 5 2.029 6 2.026 5 2.059 8 5.55 3.58

r
x=0.08 4.714 6 9.212 4 204.775 2.024 4 1.967 1
-p 2.018 0 2.027 2 2.085 7 5.94 3.81

x=0.10 4.711 7 9.206 8 204.393 1.996 7 1.963 6 2.005 2 2.067 6 2.113 4 5.16 3.77
re
x=0.12 4.710 9 9.207 7 204.338 2.001 0 1.955 2 2.004 0 2.068 0 2.118 9 5.19 3.68
lP
na
ur

2.2 Microscopic morphology


Jo

Figure 3a shows the microscopic morphology of the different doped samples at the

respective optimum sintering temperatures, where the grain and grain boundary

arrangement can be clearly observed. In addition, grain morphology presents no growth

anomalies due to Sb doping, and the average grain size sustains around 3.2 μm,

implicating that the introduction of Sb does not contribute to the change in microscopic

morphology. However, some microporosity is unavoidably created during the sintering

process, causing the samples' density to fluctuate between increasing and decreasing,

while the majority of the samples have a relative density greater than 95%, indicating

6
good sintering densification.

2.3 Raman spectra analyses

The Raman spectrogram is shown in Figure 5. A similar Raman spectrum can be found

for different doped samples, which ia ascribed to the successful entry of Sb ions into the

Ta site, maintaining the stability of the trirutile phase. Based on the group theory,

f
oo
MgTa2O6 crystals, which belong to the P42/mnm and D4h point group, have 16 Raman

r
characteristic peaks (4A1g+2B1g+4B2g+6Eg, Table S1 and Table S2). Only 11
-p
characteristic peaks can be discerned because some of them are too weak to be detected,
re
lP

as well as the overlapping of neighboring peaks with each other. The vibrational modes

corresponding to these observed peaks and the vibrational schematics are shown in
na

Figure S2 and Table S3. The characteristic peaks were separated, and frontal
ur

information was extracted from the Raman pattern using the Lorentz fitting method, as
Jo

shown in Figure 5a and Table S3. The strongest peak A1g located at 701 cm-1 is caused

by the symmetric stretching vibration of the Mg-O bond, and the peak position and the

full width at half maximum (FWHM) information contained in this peak are detached

on the right side of Figure 5a. With the increasing amount of doping, the A1g peak (v6)

shows a slight red-shift phenomenon, which is caused by the reduction of the interaction

between cations and anions due to the longer Mg-O bond and the expansion of the

MgO6 polyhedron [25, 26]. Differently from the red-shift of v6 peak, the characteristic

peaks in the middle and low Raman shift bands, caused by Ta-O bond stretching and

7
bending vibrations, such as v1, v2, and v4 peaks, show a noticeable blue-shift, as shown

in Figure 5b. The principal factor in this phenomenon is the contraction of TaO6

octahedral caused by the shortening of the bond length of the Ta-O bond induced by the

doping of Sb ions. The histogram in Figure 5c shows several typical Raman peak

intensities, normalized by the corresponding v6 peak intensities, exhibiting downward

trends depending on the degree of doping. The diminished intensity of these peaks is

f
intimately related to the polarity of the Ta-O bond, while the polarity of the Mg-O bond

oo
suffers little change due to doping. Density functional theory (DFT) was used to

r
-p
calculate the electron distribution, as shown in Figure 6(a) and 6(b). The occupation of
re
the Ta site by Sb ion leads to the aggregation of electrons at the O site, causing an
lP

increase in the Ta-O bond polarity. The increase in chemical bond polarity signifies that
na

the electron cloud near the anion is more resistant to displacement by light radiation[27],
ur

which is the primary source of the diminishing peak intensity with doping.
Jo

2.4 Electron distribution analysis

The charge density distribution of the doped crystal is calculated by the first principles

as shown in Figure 5. Obviously, compared to the charge distribution of the Ta-O bond,

the electron cloud is biased toward the O atom after the Sb ion occupies the Ta site,

which signifies a stronger ionicity between the Sb and O ions than that of the Ta-O bond.

The Mulliken charges of the pre-doped and post-doped atoms are tabulated in Table 2.

The charge of the Mg ion is close to 2, implying that Mg atoms are nearly completely

8
ionized when bonded to O atoms and that strong ionic bonds are formed between them.

In contrast, the Ta ion only loses 1.3 charges, showing a weak ionic bond between the

Ta-O bonds. Upon doping, the dopant Sb ion is ionized significantly more than the Ta

ion, delivering more charge, signaling a stronger ionic bond between Sb and O ions.

Meanwhile, lower charge loss was found for Mg and Ta ions relative to pre-doping,

which means that the ionicity of Mg-O and Ta-O bonds appear to be weakened in this

f
case. The chemical bond theory (PVL theory) [27, 28] reveals an intimate relationship

oo
between the ionicity of chemical bonds and the permittivity, as shown in the following

r
equation [34, 35].
-p
re
n2 − 1
r = +1 (1)
lP

1 − fi
Where n is the refractive index, fi is the fraction of the ionicity of the bonds. This
na

equation specifies a positive contribution of the ionicity of bonds to the permittivity, but
ur

the introduction of Sb-O with strong ionic bonds causes a weakening of the ionicity of
Jo

Ta-O and Mg-O bonds. As the number of Ta-O and Mg-O bonds is significantly greater

than the Sb-O bonds, this will entail a non-explicit monotonic trend in the total ionicity

change of the crystal. Additionally, the change in energy band structure caused by Sb

doping is shown in Figure 6a. It is found that the occupation of the Ta site by Sb atom

does not change the nature of the indirect energy band structure; rather, it causes an

additional energy band gap. To diagnose the phenomenon, the partial density of states

(PDOS) of pre-doped and post-doped crystals were calculated (Figure 6b). The valence

band maximum (VBM) is almost entirely contributed by the O s-orbital electrons, and

9
this contribution is not affected by the doping of Sb and remains frozen. On the contrary,

the contribution of the Sb s-orbital electrons to the conduction band bottom (CBB) starts

to emerge after doping. Due to the higher electronegativity, the Sb atom's s-orbital

electrons assume more energy, which causes the conduction band bottom to move

toward the higher energy region, leading to an increase in the band gap.
Table 2 Atomic Populations (Mulliken) of MgTa2O6 crystal before and after doping
Ions s p d total charge(e)

f
Before doping O1 1.88 4.87 0.00 6.75 -0.75

oo
O2 1.87 4.88 0.00 6.75 -0.75
O2’ 1.87 4.88 0.00 6.75 -0.75

r
Mg 0.29 5.87 0 6.16 1.84
Ta 0.44
-p 0.37 2.86 3.67 1.33
re
After doping O1 1.88 4.90 0.00 6.78 -0.78
O2 1.87 4.91 0.00 6.78 -0.78
lP

O2’ 1.87 4.91 0.00 6.78 -0.78


Mg 0.30 5.91 0.00 6.21 1.79
na

Ta 0.44 0.39 2.86 3.69 1.31


Sb 0.83 1.75 0.00 2.58 2.42
ur
Jo

2.5 Microwave dielectric properties

The measured permittivity variation with sintering temperature and doping is shown in

Figure 7a. Obviously, there is a significant creep up of the permittivity at the sintering

temperature stage below 1 350 °C (region Ⅰ), which is caused by the densification

process of gradual grain growth and pore exclusion of the ceramic at this stage, up to

sufficient grain growth and saturation value of the permittivity (region Ⅱ). The

measured saturation permittivity is shown as the pink histogram in Figure 7b. It was

found that the permittivity variation presented no monotonic variation with the gradient

10
doping, in contrast, it showed the same tendency as the density, which means that the

pores inside the ceramic are still the main factor affecting the permittivity as well. To

evaluate the effect of crystal structure changes, caused by Sb ion doping, on the

resulting permittivity, we first calculated the permittivity of the nonporous fully dense

ceramics using the following equation [29].

 3 p( c − 1) 
 m =  c 1 −  (2)
 2 c + 1 

f
oo
where 𝜀m , 𝜀c , and p represent the measured permittivity, the corrected permittivity, and

r
the porosity fraction, respectively. It is noteworthy that the variation in the nonporous
-p
permittivity of the doped samples did not show a monotonic shift as expected after
re
lP

eliminating the external factors. The increase in the covalence of the Mg-O and Ta-O

bonds during the doping process, as shown in Table 2, is the main cause of a
na

non-monotonic change in the dielectric constant [30].


ur
Jo

Figure 7c shows the Q×f values curve with sintering temperatures and doping. It

exhibits a temperature-dependent feature similar to the dielectric constant: The Q×f

value creeps up gradually with increasing sintering temperature; grain growth, grain

boundary fusion, and pore elimination are the main reasons for the significant increase

in Q×f value during the heating process. It is well known that the Q×f value of ceramics

is influenced by various factors, such as external factors like the second phase and the

densities of the samples, as well as internal factors like the crystal structure and

crystallinity. In this work, a single-phase compound was synthesized, and the density

11
became the main external factor affecting the Q×f value. The relationship between the

crystal structure and the Q×f value can be expressed by the atomic packing fraction (fp),

as in the following equation[7].

volume of packed ions


fp = Z (3)
volume of unit cell
Where Z is 2 for MgTa2O6 structure, the calculated results are listed in Table 3. It can be

found that when the doping amount is less than 0.08, the fp value and the relative

f
oo
density of the sample increase with the increase in the doping amount. The Q×f value,

r
on the other hand, shows a slight decrease, which is due to the decrease in crystallinity
-p
after doping, as shown by the broadening Raman peak in Figure 4a [31, 32]. With the
re
lP

doping amount above 0.08, the crystallinity and the relative density of the sample both

decrease, which leads to a significant decrease in the Q×f value.


na

Table 3 Q×f values, the fp values, Raman FWHM, and the relative density of Sb-doped MgTa2O6
ceramics.
ur

x Q×f fp (%) FWHM Relative


(104GHz) density (%)
Jo

0.02 12.72 70.93 20.44 94.0


0.04 12.44 70.92 20.41 95.4
0.06 12.34 70.99 20.84 96.2
0.08 11.20 71.00 21.51 96.5
0.10 10.59 71.12 21.37 95.4
0.12 9.75 71.13 22.15 94.5

The temperature coefficient of resonant frequency (TCF, τf) is an attractive parameter

for application as an evaluation index of ceramics’ thermal stability. Being defined as

the offset of the operating frequency at different temperatures, its value is closer to zero,

12
indicating the higher stability of the material. Figure 7d shows the variation of TCF

values with different doping, and a significant decrease in TCF values with increasing

doping can be observed. Generally, the TCF derives mainly from the change of the

dielectric (τε) and linear thermal expansion coefficient (αL) under heating [33, 34]:


 f = −(  +  )L (4)
2

where αL maintains at 14 in the present system (see attached calculation). From this, the

f
oo
variation of τε is the key contributing factor to the variation of TCF, which can be given

r
by the following equation [35, 36]: -p
( -1)( + 2)  1 d m 
re
 =  − 3 L  (5)
3   m dT 
lP

Where ε is the permittivity, and the αm is the polarizability. It can be found that τε is
na

negatively correlated with the polarizability. The typical discrepancy of polarizability in


ur

different doped molecules is the main factor of τε change. Doping causes a decrease in
Jo

the polarizability of the molecule, resulting in an increase in for different doped

ceramics because the polarizability of Sb (4.27 Å3) is lower than that of Ta (4.73 Å3).

Also, equation 4 clarifies that the change in TCF and the change in dielectric are

negatively correlated, from which it can be concluded that the decrease in polarizability

caused by Sb doping is the key factor leading to the decrease in TCF from 36 to 15×10-6

o
C-1.

4. Conclusion

13
In this study, Sb ions managed to enter the lattice of MgTa2O6 and occupy Ta sites to

form a pure trirutile structure. Experimentally, doping was found to cause contraction of

TaO6 polyhedral and inflation of MgO6 octahedral, which were responsible for the

blue-shift of the low frequency band peaks and the red-shift of the high frequency band

peaks in Raman. The first principle calculation revealed that the occupation of the Ta

site by Sb ion provided more electrons to the anion, which conversely resulted in a

f
reduction of the original Ta and Mg ionization. The microwave dielectric properties of

oo
the doped samples all exhibited robust temperature-dependent behavior. However, the

r
-p
measured dielectric constants did not show a strong doping dependence. Doping of Sb
re
also causes a decrease in the crystallinity of the crystals, resulting in a slight decrease in
lP

the quality factor associated with the doping. In conclusion, Sb doping improves the
na

sintering characteristics and thermal stability of the original MgTa2O6 ceramics while
ur

maintaining the stability of the physical phase, and only at the slight expense of the
Jo

dielectric constant and quality factors.

Acknowledgments

This work was supported by National Key Scientific Instrument and Equipment

Development Project (No.51827802), and National Natural Science Foundation of

China (No.51902042).

References
[1] G. Wang, D. Zhang, X. Huang, Y. Rao, Y. Yang, G. Gan, Y. Lai, F. Xu, J. Li, Y. Liao, C. Liu, L. Jin, V.G. Harris,
H. Zhang, Crystal structure and enhanced microwave dielectric properties of Ta5+ substituted Li3Mg2NbO6
ceramics, J. Am. Ceram. Soc. 103 (2019) 214-223.

14
[2] W. Wersing, Microwave ceramics for resonators and filters, Curr. Opin. Solid. St. & M. 1 (1996)
715-731.
[3] I.M. Reaney, D. Iddles, Microwave Dielectric Ceramics for Resonators and Filters in Mobile, J. Am.
Ceram. Soc. 89 (2006) 2063-2072.
[4] T. A.Vanderah., Talking Ceramics, Sci. 298 (5596) (2002) 1182-1184.
[5] L. Ni, L. Li, M. Du, Ultra-high-Q and wide temperature stable Ba(Mg1/3Ta )O3 microwave dielectric
ceramic for 5G-oriented dielectric duplexer adhibition, J. Alloy. Compd. 844 (2020).
[6] R.G.M. Oliveira, R.A. Silva, J.E.V. de Morais, G.S. Batista, M.A.S. Silva, J.C. Goes, H.D. de Andrade, I.S.
Queiroz Júnior, C. Singh, A.S.B. Sombra, Effects of CaTiO3 addition on the microwave dielectric properties
and antenna properties of BiVO4 ceramics, Composites Part B: Engineering 175 (2019).
[7] P. Wang, Y.R. Wang, J.X. Bi, H.T. Wu, Effects of Zn2+ substitution on the crystal structure, Raman
spectra, bond energy and microwave dielectric properties of Li2MgTiO4 ceramics, J. Alloy. Compd. 721

f
(2017) 143-148.

oo
[8] A.C.G. R. C. Kell, G. C. E. Olds, High-Permittivity Temperature-Stable Ceramic Dielectrics with Low
Microwave Loss, J. Am. Ceram. Soc 56 (1973) 352-354.

r
[9] M.S. Fu, X.Q. Liu, X.M. Chen, Structure and microwave dielectric characteristics of Ca1−xNd2x/3TiO3
-p
ceramics, J. Eur. Ceram. Soc. 28(3) (2008) 585-590.
[10] Hyo-Jong Lee, I.-T.K.a.K.S. Hong, Dielectric properties of AB2O6 compounds at microwave
re
frequencies (A=Ca, Mg, Mn, Co, Ni, Zn,and B=Nb, Ta), J. Appl. Phys. 36 (1997) 1318-1320.
[11] M. Dang, H. Ren, X. Yao, H. Peng, T. Xie, H. Lin, L. Luo, Investigation of phase composition and
lP

microwave dielectric properties of MgO-Ta2O5 ceramics with ultrahigh Qf value, J. Am. Ceram. Soc. 101(7)
(2018) 3026-3031.
na

[12] W.C. Tzou, Y.C. Chen, C.F. Yang, C.M. Cheng, Microwave dielectric characteristics of Mg(Ta1−xNbx)2O6
ceramics, Mater. Res. Bull. 41(7) (2006) 1357-1363.
[13] E.S. Kim, S.J. Kim, Effect of Sintering Temperature on the Microwave Dielectric Properties of
ur

MgB2O6(B = Nb5+, Ta5+) Ceramics, Ferroelectrics 388(1) (2009) 93-100.


Jo

[14] H. Rao, S. Ye, W. Sun, W. Yan, Y. Li, H. Peng, Z. Liu, Z. Bian, Y. Li, C. Huang, A 19.0% efficiency achieved
in CuOx-based inverted CH3NH3PbI3−xClx solar cells by an effective Cl doping method, Nano Energy 27
(2016) 51-57.
[15] H. Ding, L. Xu, C. Wen, J.J. Zhou, K. Li, P. Zhang, L. Wang, W. Wang, W. Wang, X. Xu, W. Ji, Y. Yang, L.
Chen, Surface and interface engineering of MoNi alloy nanograins bound to Mo-doped NiO nanosheets
on 3D graphene foam for high-efficiency water splitting catalysis, Chem. Eng. J. s440 (2022) 135847.
[16] C.L. Huang, K.H. Chiang, C.Y. Huang, Microwave dielectric properties and microstructures of MgTa2O6
ceramics with CuO addition, Mater. Chem. Phys. 90(2-3) (2005) 373-377.
[17] C. Huang, J. Chen, Synthesis, Crystal Structure, and Microwave Dielectric Properties of
(Mg1−xCox)Ta2O6 Solid Solutions, J. Am. Ceram. Soc. 93(2) (2010) 470-473.
[18] C.L. Huang, J.Y. Chen, Y.W. Tseng, C.Y. Jiang, G.S. Huang, High Dielectric Constant and Low-Loss
Microwave Dielectric Ceramics Using (Zn0.95M2+0.05)Ta2O6 (M2+=Mn, Mg, and Ni) Solid Solutions, J. Am.
Ceram. Soc. 93(10) (2010) 3299-3304.
[19] L. Shi, X. Wang, R. Peng, G. Wang, C. Liu, X. Shi, D. Zhang, H. Zhang, Crystallographic characteristics
and microwave dielectric properties of Ni-modified MgTa2O6 ceramics, Ceram. Int. 47(16) (2021)
22514-22521.
[20] L. Shi, D. Zhang, R. Peng, C. Liu, X. Shi, X. Wang, H. Zhang, Investigation of crystal characteristics,

15
Raman spectra, and microwave dielectric properties of Mg1-xZnxTa2O6 ceramics, J. Eur. Ceram. Soc. 41(11)
(2021) 5526-5530.
[21] Y. Zhang, Z. Yue, X. Qi, B. Li, Z. Gui, L. Li, Microwave dielectric properties of Zn(Nb1−xTax)2O6 ceramics,
Mater. Lett. 58(7-8) (2004) 1392-1395.
[22] M. Xiao, S. He, J. Lou, P. Zhang, Structure and microwave dielectric properties of MgZr(Nb1−xSbx)2O8
(0≤x≤0.1) ceramics, J. Alloy. Compd. 777 (2019) 350-357.
[23] X. Guo, N. Zhu, M. Xiao, X. Wu, Structural and Dielectric Properties of Ag(Nb0.8Ta0.2)1-xSbxO3 (x<0.08)
Ceramics, J. Am. Ceram. Soc. 90(8) (2007) 2467-2471.
[24] Y.R. Luo, Comprehensive handbook of chemical bond energies, CRC press2007.
[25] L. Cao, Y. Yuan, E. Li, S. Zhang, Relaxor regulation and improvement of energy storage properties of
Sr2NaNb5O15-based tungsten bronze ceramics through B-site substitution, Chem. Eng. J. 421 (2021).
[26] Franklin D. Hardcastle, I.E. Wachs, Determination of Vanadium-Oxygen Bond Distances and Bond

f
Orders by Raman, J. phys. Chem. 95 (1991) 5031-5041.

oo
[27] V. Wagner, Liang, J.J., Kruse, R., Gundel, S., Keim, M., Waag, A. and Geurts, J., Lattice Dynamics and
Bond Polarity of Be-Chalcogenides A New Class of II–VI Materials, phys. stat. sol. 215 (1999) 87-91.

r
[28] Huanrong Tian, Jinjie Zheng, Lintao Liu, Haitao Wu, Hideo Kimura, Yizhong Lu, Z. Yue., Structure
-p
characteristics and microwave dielectric properties of Pr2(Zr1-xTix)3(MoO4)9 solid solution ceramic with a
stable temperature coefficient, J. Mater. Sci. & Tech. 116 (2022) 121-129.
re
[29] S.J. Penn, N.M. Alford, A. Templeton, X.R. Wang, M.S. Xu, M. Reece, K. Schrapel, Effect of porosity
and grain size on the microwave dielectric properties of sintered alumina, J. Am. Ceram. Soc. 80(7) (1997)
lP

1885-1888.
[30] H. Yang, S. Zhang, H. Yang, Y. Yuan, E. Li, Influence of (Al1/3W2/3)5+ co-substitution for Nb5+ in NdNbO4
na

and the impact on the crystal structure and microwave dielectric properties, Dalton Trans 47(44) (2018)
15808-15815.
[31] Q. Liao, L. Li, Structural dependence of microwave dielectric properties of ixiolite structured
ur

ZnTiNb2O8 materials: crystal structure refinement and Raman spectra study, Dalt. Trans. 41(23) (2012)
Jo

6963-9.
[32] J. Zhang, R. Zuo, A. Feteira, Effect of Ordering on the Microwave Dielectric Properties of
Spinel-Structured (Zn1−x(Li2/3Ti1/3)x)2TiO4 Ceramics, J. Am. Ceram. Soc. 99(10) (2016) 3343-3349.
[33] P.L. Wise, I.M. Reaney, W.E. Lee., Tunability of τf in perovskites and related compounds, J. Mater. Res.
17 (2002) 2033-2040.
[34] G. Wang, J. Li, X. Zhang, Z. Fan, F. Yang, A. Feteira, D. Zhou, D.C. Sinclair, T. Ma, X. Tan, D. Wang, I.M.
Reaney, Ultrahigh energy storage density lead-free multilayers by controlled electrical homogeneity,
Energy Environ. Sci. 12(2) (2019) 582-588.
[35] E.L. Colla, I.M. Reaney, N. Setter, Effect of structural changes in complex perovskites on the
temperature coefficient of the relative permittivity, J. Appl. Phys. 74(5) (1993) 3414-3425.
[36] Z. Lu, H. Zhang, W. Lei, D.C. Sinclair, I.M. Reaney, High-Figure-of-Merit Thermoelectric La-Doped
A-Site-Deficient SrTiO3 Ceramics, Chem. Mater. 28(3) (2016) 925-935.

Figure caption

16
Figure 1 XRD patterns of Sb-doped MgTa2O6 ceramics with different x content.

Figure 2 Rietveld refinement results; a: Refinement patterns of Mg(Ta0.98Sb0.02)2O6

ceramics; b: Variation of lattice parameters (a, b, and V) with doping; c: Schematic

diagram of MgTa2O6 crystal structure and TaO6 and MgO6 octahedral.

Figure 3 SEM images of Sb-doped MgTa2O6 ceramics with different x content (a, scale

bar is 2 μm); The curves of density and relative density of the samples with doping at

f
the optimum sintering temperature (b); And the statistical plots of grain size distribution

oo
and average grain size variation for different doped samples (c).

r
-p
Figure 4 Raman spectra and the deconvoluted Raman vibrational peak of doped
re
MgTa2O6 ceramics (a), the right one is a partial enlargement of the light purple area;
lP

The offset of the peak position of several characteristic peaks relative to the 0.02-doped
na

characteristic peak (b, increase indicates a blue-shift and decrease indicates a red-shift);
ur

The histogram of the relative intensity of several characteristic peaks relative to v6 with
Jo

doping.

Figure 5 Three-dimensional distribution of charge density (a) and sliced charge density

isopotential graph (b) of Sb-doped MgTa2O6 crystal.

Figure 6 Energy band structure and partial density of states (PDOS) diagrams of

pre-doping and post-doping crystal.

Figure 7 Diagram of microwave dielectric properties of doped samples with the

variation curve of dielectric constant with sintering temperature (a); histogram of

saturation dielectric constant and pore-free dielectric constant with doping (b); diagram

17
of Q×f values with sintering temperature (c) and saturation Q×f values with doping

(inset); diagram of TCF value with doping (d).

f
r oo
-p
re
lP
na
ur
Jo

18
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Highlights
⚫ The entry of Sb5+ into the lattice causes shrinkage of the lattice and TaO6
octahedra, and dilation of the MgO6 polyhedra.
⚫ DFT calculations show that Sb ionizes more electrons when bonded to oxygen,
while Mg and Ta ionization diminishes.
⚫ Reduce τf value to 15×10-6 oC-1 while maintaining εr ~27, and high Q×f value of
109 000 GHz for 0.08 Sb-doped ceramics.

f
r oo
-p
re
lP
na
ur
Jo
Liang Shi is a PhD candidate of School of Electronic Science and Engineering,

University of Electronic Science and Technology of China, Chengdu, China. He

obtained his B.S. in 2015 from University of Electronic Science and Technology of

China. His research interests include low temperature co-fired ceramics, microwave

dielectric materials and perovskite thin film.

Huaiwu Zhang is a professor in the School of Electronic Science and Engineering,

f
oo
University of Electronic Science and Technology of China, Chengdu, China. His

r
-p
research interests include magneto-electric information materials and devices,
re
magnetic nanomaterial chips, RF and microwave materials and devices, integrated
lP

LTCC/LTCF devices.
na
ur
Jo
Liang Shi

f
oo
Huaiwu Zhang

r
-p
re
lP
na
ur
Jo
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.

f
r oo
-p
re
lP
na
ur
Jo

You might also like