Next Article in Journal
Simulation of a New CZTS Solar Cell Model with ZnO/CdS Core-Shell Nanowires for High Efficiency
Previous Article in Journal
Flux Method Growth and Structure and Properties Characterization of Rare-Earth Iron Oxides Lu1−xScxFeO3 Single Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Li+ Ionic Conduction and Relaxation Properties of Li5+2xLa3Ta2-xGaxO12 Garnets

1
Department of Physics, College of Science, King Faisal University, Hofuf 31982, Saudi Arabia
2
Department of Physics, Faculty of Science, The New Valley University, El-Kharga 72511, Egypt
3
Physics Department, Faculty of Science, Assiut University, Assiut 71516, Egypt
4
Department of Basic Engineering Sciences, College of Engineering, Imam Abdulrahman Bin Faisal University, Dammam 34221, Saudi Arabia
5
Department of Physics, University of Petroleum and Energy Studies, Dehradun 248007, India
6
Department of Applied Molecular Chemistry, College of Industrial Technology, Nihon University, Narashino 275-8575, Japan
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(6), 770; https://doi.org/10.3390/cryst12060770
Submission received: 14 April 2022 / Revised: 19 May 2022 / Accepted: 23 May 2022 / Published: 26 May 2022
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
In the current work, we studied the effect of Ga+3 substitutions on the Ta+5 sites in Li5+2xLa3Ta2-xGaxO12 (LLT-Ga) lithium conducting garnets (with x = 0.1–0.5) in order to enhance the ionic conductivity of these materials. The current materials are prepared by solid state reaction and their electrical properties are studied by impedance spectroscopy measurements. XRD data showed that cubic garnet phases are obtained for LLT-Ga garnets. The ionic conductivity increased by one order of magnitude for x = 0.3 composition with a value of ~4 × 10−5 S/cm compared to that of Li5La3Ta2O12 material. Moreover, the hopping frequency and the concentration of mobile Li+ ions were estimated from analysis of the conductivity spectra, and it was found that both the concentration and mobility of Li+ ions increased with increasing Ga+3 content in the materials. The dielectric and relaxation properties were studied in the dielectric permittivity and electric modulus formalisms. The current materials exhibited giant values of the dielectric constant of ε′ ~ 6500, originating from internal effects in the materials.

1. Introduction

Lithium-ion batteries are currently major power sources in our daily lives, where they are used in various electronic devices. [1,2]. Due to safety issues arising from organic electrolytes in modern batteries, there are appeals for the development of solid-state batteries using solid inorganic electrolytes. [3,4]. Several materials with good lithium ionic conductivity have been studied and lithium conducting garnet materials, Li5La3M2O12 (LLM, M = Ta or Nb) and Li7La3Zr2O12 (LLZ), are considered interesting materials due to their promising electrical properties [4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27]. LLM garnets have ionic conductivity of 10−6 S/cm at RT, whereas the ionic conductivity of LLZ reaches 10−4 S/cm at RT [4,5,6,7,8,9,10,11,12,13,14,15,16,18]. These garnet materials are purely ionic with negligible electronic conductivity [4,5,6,18]. Due to these promising properties LLZ garnets have attracted much research interest for possible applications as electrolytes in all solid-state batteries. LLZ could have a cubic (highly conductive) or tetragonal (poorly conductive) phase. The cubic phase is stabilized by high temperature sintering above 1200 °C. However, this high temperature sintering is usually associated with Al contamination from the alumina crucibles. It is found that Al-substitution on Li sites is the reason for the stabilization of the cubic phase of LLZ. The ionic conductivity of LLZ could be enhanced by chemical substitutions, either in the Li sites, by AL+3, Ga+3, Y+3, or in the Zr sites, by higher valency cations, such as Ta+5, Nb+5, Te+6 and W+6, leading to ionic conductivity >10−3 S/cm [18,19,20,21,22,23,24,25]. The ionic conductivity of 10−6 S/cm at RT of LLT is relatively low, leading to extensive research efforts to improve the ionic conductivity of these materials. Chemical substitutions by divalent/trivalent cations on the La+3/Ta+5, such as in Li6ALa2Ta2O12, A = Ba, Ca, Sr, Mg, [2,3,4,5,6,7,8,9] and in Li5+2xLa3Ta2-xYxO12 [13], lead to enhanced ionic conductivity to 10−5–10−4 S/cm at RT [5,6,7,8,9,10,11,12,13]. Other factors could also influence the ionic conductivity of garnet materials including the preparation techniques and processing conditions, such as the calcinations temperature and sintering steps [4,5,6,7,8,9,10,11,12,13,14]. Chemical substitution by divalent cations on the La sites, or trivalent cations on the Ta sites, in LLT garnets, will increase the lithium content in the materials, leading to increasing ionic conductivity.
Until now, there are no studies on the effect of Ga+3 substitutions on ionic conduction in LLT garnets. Moreover, most of the research work on ionic conduction in garnets is performed at ambient and high temperatures, and few low temperature studies have been reported. Low temperature measurements are expected to help in understanding the conduction and relaxation properties of garnet materials. Moreover, the presence of high-density mobile Li+ ions could induce dielectric dipoles, leading to high values of the dielectric constant. Therefore, the aims of the present work are: (i) synthesis of Li5+2xLa3Ta2-xGaxO12 (LLT-Ga; x = 0.1–0.5) lithium conducting garnets in order to enhance lithium ionic conductivity, (ii) study of the relaxation dynamics in garnet materials through analysis of the electric modulus spectra at different temperatures, and (iii) exploring the dielectric properties of LLT-Ga garnets.

2. Materials and Methods

Li5+2xLa3Ta2-xGaxO12 (with x = 0.1, 0.2, 0.3, 0.4, 0.5; all the samples LLT-Ga1, LLT-Ga2, LLT-Ga3, LLT-Ga4 and LLT-Ga5, respectively) have been synthesized by mechanical milling and solid-state reaction techniques. High purity Li2CO3, La2O3, Ta2O3 and Ga2O3 were used. Excess 10 wt% of Li2CO3 was added for compensation of Li2O. The mixed powder was ball milled for 12 h with 250 rpm. The resulting powder was calcined at 700 °C for 12 h, then ball milled again under the same conditions. A second calcination step was performed at 900 °C for 12 h followed by final milling at 400 rpm for 3 h. The ball milling steps were performed in 2-propanol using Fritsch P-7 premium line machine with 45 mL pots and 14 balls of 10 mm diameters, both made from tungsten carbide. The balls to powder mass ratio was 8:1. The obtained powder was pressed under 2.5 ton in a 10 mm diameter die to get pellets with a 1–2 mm thickness. The sintering of the samples was performed at 1000 °C for 12 h in alumina crucibles. To prevent possible Al contamination of the samples during the sintering process in alumina crucibles the pellets were buried in a large quantity of the powder to prevent direct contact of the pellets with the alumina crucibles. Structural characterization of LLT-Ga samples was studied by XRD measurements in the 10 ≤ 2θ ≤ 80 range using monochromated radiation (λ = 1.5406 Å) with a step of 0.02° using a Bruker D8 advance diffractometer. SEM measurements were performed with a field emission scanning electron microscope, Joel SM7600F, to determine the grain size of the materials. The electrical and dielectric properties of LLT-Ga samples were studied by impedance spectroscopy. Silver paste was applied on both surfaces of the sintered pellets as electrodes. The impedance data were collected in the 1–107 Hz frequency range at different temperatures using the Novocontrol concept 50 system.

3. Results and Discussion

The XRD patterns of LLT-Ga samples are shown in Figure 1. The observed patterns in Figure 1 agree with the standard XRD database of Li5La3Ta2O12 (PDF#45-0109), indicating the formation of the cubic garnet structure. Minor impurity peaks are also observed in the XRD patterns of these materials. The Rietveld refinement method was deployed using RIETAN-2000 software [28] for the investigated materials in order to elucidate the lattice parameters of the materials.
The results in Figure 1 indicate that LLT-Ga samples have cubic structure with the space group I a 3 ¯ d with an average lattice constant of 12.827 Å for x = 0.1–0.5 samples. The value of the lattice constant of the current LLT-Ga garnets agrees well with the value of 12.85 Å reported previously for LLT samples prepared by the sol-gel technique [14]. Moreover, we noticed that there was no change in the lattice constant with increasing Ga+3 content due to the close matching between the ionic radius of Ga+3 (0.62 Å) with that of Ta+5 ions (0.64 Å). In Figure 1 we notice the presence of impurity peaks in the XRD patterns. One major impurity peak was observed in x = 0.1 sample at 23.8°, whereas the impurity content increased for x = 0.5 sample, as evidenced by the presence of several impurity peaks at 23.5°, 29.9°, 41.2°, 42.5° and 56.2°. The impurity phases could be assigned to LiGaO2 (PDF# 72-1640) and La2O3 (PDF# 74-1144). SEM micrographs for LLT-Ga ceramics are shown in Figure 2. Dense ceramics with well-defined coarse grains were obtained for all Ga+3 compositions, with grain size in the 3–5 μm range. It can be observed in Figure 2 that no noticeable changes in the grain size or the morphology of the ceramics were observed with increasing Ga+3 substitutions in the materials.
Impedance diagrams of LLT-Ga3 ceramics are shown in Figure 3 as a representative example. A single semicircle at the high frequency region is observed, where the effect of the grain boundaries could not be distinguished in this figure. Therefore, the real axis intercept of the semicircles represents the total resistance, R, of the materials. The total ionic conductivity was calculated by the relation; σ d c = l / R A , where l is the sample thickness and A is the surface area. The variation of the lithium ionic conductivity of LLT-Ga ceramics with temperature is depicted in Figure 4, where the conductivity follows the Arrhenius relation;
σ d c = σ o   exp ( Δ E k T )
The values of σ d c at 300 K of Li5+2xLa3Ta2-xGaxO12 ceramics are summarized in Table 1. It is clear that the Li5.6La3Ta1.7Ga0.3O12 garnets exhibited the highest conductivity with a value of 4 × 10−5 S/cm compared to 4.26 × 10−6 S/cm of pure Li5La3Ta2O12, indicating considerable lithium conductivity enhancement is achieved by Ga+3 substitutions in LLT garnets. It is clear from Figure 4 and the data in Table 1 that the most effective doping region was up to x = 0.1 of Ga+3 content, where for larger Ga+3 content, between x = 0.2–0.4, the conductivity was almost saturated. On further increase in Ga+3 content (i.e., for x = 0.5) conductivity dropped slightly, due to the increased content of impurity phases in this composition. The activation energy of conduction equaled 0.55 eV for all LLT-Ga samples, which is comparable to that of pure LLT garnets of 0.55–0.60 eV [5,14].
According to the chemical formula Li5+2xLa3Ta2-xGaxO12 of Ga+3 substituted garnets the Li+ content per unit formula was expected to increase from 5Li in pure LLT to (5 + 2x)Li for x content of Ga+3. The increased content of Li+ ions could be the reason for the enhanced conductivity of Ga substituted LLT materials. Furthermore, Ga+3 substitutions in LLT-Ga materials may create accessible vacant sites for Li+ ions, which enhance mobility and conductivity. In the following, we will estimate the hopping frequency (ωH) and the concentration (nc) of mobile Li+ ions in LLT-Ga garnets. The conductivity spectra of LLT-Ga3 ceramics are shown in Figure 5. A frequency independent dc conductivity was observed at low temperatures and frequencies, followed by a dispersion region where σ ( ω ) increased with frequency at the high frequency range. The conductivity data were analyzed by the following formula [29,30,31,32],
σ ( ω ) = σ d c   [ 1 + ( ω ω H ) n ]
and the dc conductivity in Equation (2) could be expressed as:
σ d c = e n c μ = n c e 2 γ λ 2 k T ω H
In the above equations n represents the exponent of the dispersion region of the conductivity, γ = 1/6 is a geometrical factor for 3D ion hopping and λ is the hopping distance. The fitting results of the conductivity spectra are shown in Figure 5, and the temperature variation of the estimated parameters σdc and ωH is shown in Figure 6. The values of the activation energy for the conduction, Eσ, and the hopping process, EH, were almost the same [see Table 1]. The values of nc for mobile Li+ ions in LLT-Ga ceramics were estimated [see Table 1] using Equation (3) with a value of λ = 1.7 Å [33]. The values of nc increased by a factor of ~1.8 after Ga substitutions in LLT-Ga garnets. On the other hand, the hopping frequency of Li+ ions increased considerably as Ga+3 content increased up to x = 0.3 [see Figure 6]. These observations suggest that enhanced Li+ ionic conductivity in LLT-Ga ceramics is due to the increased concentration and mobility of Li+ ions up to x = 0.3 of Ga+3 content. For x = 0.4 and 0.5 compositions, the hopping frequency decreased leading to a decrease in ionic conductivity. This drop in the hopping frequency could be attributed to the increased content of impurity phases in the materials.
In garnet structure Li+ ions could occupy either the immobile 24 d tetrahedral Li(1) sites or the mobile 48 g/96 h octahedral Li(2) sites [15,16]. The occupancy of Li+ ions in these sites depends on the total Li+ content per unit formula in the material. If Li+ content is low (3Li in Li3Nd3Te2O12 for example), Li+ ions will occupy all the immobile Li(1) sites and the Li(2) sites are kept empty, leading to poor ionic conductivity. For garnet materials with higher Li+ content, such as in LLT (5Li per unit formula), our current materials LLT-Ga [(5 + 2x)Li per unit formula] or LLZ (7Li per unit formula), redistribution of Li+ occupancy occurs where the occupancy of the tetrahedral Li(1) sites decreases and occupancy of the octahedral Li(2) sites increases [15,16]. This redistribution process leads to more vacant Li(1) sites which become accessible for the movement of Li+ ions and facilitate the 3D diffusion pathway of Li+ through 24 d-96 h-48 g-96 h-24 d chain [26,27]. Therefore, the observed enhancement of the ionic conductivity in LLT-Ga is due to both increasing concentration and hopping frequency (i.e., the mobility) of mobile Li+ ions with Ga+3 substitution in the materials.
Different classes of materials, including ferroelectrics, perovskite oxide dielectrics and ionic conductors, exhibit giant dielectric constant [34,35,36,37,38]. The frequency dependence of the dielectric constant ε′ for Li5.2La3Ta1.9Ga0.1O12 garnets is shown in Figure 7a. A plateau region with ε′ ~ 6500 was observed at low frequencies, whereas the high frequency plateau represented the bulk effect with ε′ ~ 35. Nyquist plots of the dielectric data at 280 K are shown in Figure 7b. In this figure, three contributions can be distinguished. At high frequencies the value of the dielectric constant was ε′ ~ 35 [inset of Figure 7b], which corresponded to the grain dielectric permittivity. A large semicircle with a value of ε′ ~ 6500 was observed in the intermediate frequency region followed by a spike at lower frequencies. The observed spike in Figure 7b could be due to the electrode polarization effect at the sample-electrode interface region. Therefore, the high value of ε′ ~ 6500 could originate from other internal effects in the materials, such as the formation of dipolar polarization due to the displacement of Li+ ions in the octahedral, similar to Li5+2xLa3Ta2-xYxO12 [13]. ε′ spectra of all LLT-Ga ceramics at 300 K are shown in Figure 7c for comparison. All of the investigated ceramics exhibited similar dielectric response with giant dielectric constant values. The spectra of tan δ (tan δ = ε″/ε′) for x = 0.10 composition is shown in Figure 8 at selected temperatures. The relaxation time, calculated from the peak frequency of tan δ spectra, and its temperature dependence are shown in Figure 8. The values of the activation energy of the dielectric relaxation process are summarized in Table 1, which agree with activation of the ionic conduction process.

4. Conclusions

In the present work we successfully synthesized Li5+2xLa3Ta2-xGaxO12 lithium conducting garnets with cubic structure. Impurity phases became more prominent for higher content of Ga+3 ≥ 0.40. The ionic conductivity of the Ga substituted garnets was one order of magnitude larger than that of pure LLT materials, with x = 0.30 composition exhibiting the highest conductivity of 4–10−5 S/cm at 300 K. We estimated the values of the hopping frequency and the concentration of mobile Li+ ions in the investigated materials. It was concluded that the enhanced ionic conductivity in LLT-Ga garnets is due to increasing both the concentration and mobility of Li+ ions in the materials, with mobility having a more prominent role. The relaxation properties of the investigated materials were studied through dielectric permittivity formalisms. The dielectric relaxation time from the dissipation factor was determined and found to be thermally activated with the same activation energy of the conduction process. This result indicates that Li+ ions are responsible for long-range transport as well as the local re-orientation relaxation in garnet materials. The current materials exhibited giant dielectric values of ε′ ~ 6500, which were not related to surface polarization at the sample/electrode interface.

Author Contributions

Conceptualization, M.M.A.; methodology, M.M.A., F.R.A.-G. and H.M.K.; validation, M.M.A. and H.M.K.; formal analysis, M.M.A., F.R.A.-G., H.M.K., S.K. and K.Y.; investigation, M.M.A., F.R.A.-G., H.M.K., S.A.A., T.S.K. and H.A.K.; resources, M.M.A. and H.M.K.; data curation, M.M.A., H.M.K. and K.Y.; writing—original draft preparation, M.M.A.; writing—review and editing, H.A.K. and S.A.A.; supervision, M.M.A. and H.M.K.; project administration, H.M.K.; funding acquisition, H.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research in King Faisal University (Saudi Arabia), grant number 1811017 and the APC was funded by the same grant number 1811017.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Acknowledgments

The authors acknowledge the Deanship of Scientific Research at King Faisal University (Saudi Arabia) for financial support under the Research Group Support Track (grant number 1811017).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armand, M.; Tarascon, J.-M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  2. Hasoun, J.; Panero, S.; Reale, P.; Scrosati, B. A new, safe, high-rate and high-energy polymer lithium-ion battery. Adv. Mater. 2009, 21, 4807–4810. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, C.H.; Xie, S.; Sperling, E.; Yang, A.S.; Henriksen, G.; Amine, K. Stable lithium-ion conducting perovskite lithium-stronsium-tantalum-zirconium-oxide system. Solid State Ion. 2004, 167, 263–272. [Google Scholar] [CrossRef]
  4. Thangadurai, V.; Weppner, W. Li6ALa2Ta2O12 (A = Sr, Ba): Novel garnet-like oxides for fast lithium ion conduction. Adv. Funct. Mater. 2005, 15, 107–112. [Google Scholar] [CrossRef]
  5. Thangadurai, V.; Kaack, H.; Weppner, W. Novel fast lithium ion conduction in garnet-type Li5La3M2O12 (M. = Nb, Ta). J. Am. Ceram. Soc. 2003, 86, 437–440. [Google Scholar] [CrossRef]
  6. Thangadurai, V.; Weppner, W. Investigations on electrical conductivity and chemical compatibility between fast lithium ion conducting garnet-like Li6BaLa2Ta2O12 and lithium battery cathodes. J. Power Sources 2005, 142, 339–344. [Google Scholar] [CrossRef]
  7. Murugan, R.; Thangadurai, V.; Weppner, W. Effect of lithium ion content on the lithium ion conductivity of the garnet-like structure Li5+xBaLa2Ta2O11.5+0.5x (x = 0 − 2). Appl. Phys. A 2008, 91, 615–620. [Google Scholar] [CrossRef]
  8. Murugan, R.; Thangaduraiand, V.; Weppner, W. Lithium ion conductivity of Li5+xBaxLa3-xTa2O12 (x = 0 − 2) with garnet-related structure in dependence of the barium content. Ionics 2007, 13, 195–203. [Google Scholar] [CrossRef]
  9. Murugan, R.; Thangaduraiand, V.; Weppner, W. Lattice parameter and sintering temperature dependence of bulk and grain-boundary conduction of garnet-like solid Li-electrolytes. J. Electrochem. Soc. 2008, 155, A90–A101. [Google Scholar] [CrossRef]
  10. Awaka, J.; Kijima, N.; Takahashi, Y.; Hayakawa, H.; Akimoto, J. Synthesis and crystallographic studies of garnet-related lithium-ion conductors Li6CaLa2Ta2O12 and Li6BaLa2Ta2O12. Solid State Ion. 2009, 180, 602–606. [Google Scholar] [CrossRef]
  11. Zhong, Y.; Zhou, Q.; Guo, Y.; Li, Z.; Qiang, Y. The ionic conductivity of Li6BaLa2M2O12 with coexisting Nb and Ta on the M sites. Ionics 2013, 19, 697–700. [Google Scholar] [CrossRef]
  12. Kokal, I.; Ramanujachary, K.V.; Notten, P.H.L.; Hintzen, H.T. Sol-gel synthesis and lithium ion conduction properties of garnet-type Li6BaLa2Ta2O12. Mater. Res. Bull. 2012, 47, 1932–1935. [Google Scholar] [CrossRef]
  13. Baral, A.K.; Narayanan, S.; Ramezanipour, F.; Thangadurai, V. Evaluation of fundamental transport properties of Li-excess garnet-type Li5+2xLa3Ta2-xYxO12 (x = 0.25, 0.5 and 0.75) electrolytes using ac impedance and dielectric spectroscopy. Phys. Chem. Chem. Phys. 2014, 16, 11356–11365. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, Y.X.; Wang, X.P.; Wang, W.G.; Fang, Q.F. Sol-gel synthesis and electrical properties Li5La3Ta2O12 lithium ionic conductors. Solid State Ion. 2010, 181, 33–36. [Google Scholar] [CrossRef]
  15. Cussen, E.J. The structure of lithium garnets: Cation disorder and clustering in a new family of fast Li+ ion conductors. Chem. Commun. 2006, 4, 412–413. [Google Scholar] [CrossRef] [Green Version]
  16. O’Challaghan, M.P.; Cussen, E.J. Lithium dimer formation in the Li-conducting garnets Li5+xBaxLa3-xTa2O12 (0 < x ≤ 1.6). Chem. Commun. 2007, 20, 2048–2050. [Google Scholar]
  17. Xu, M.; Park, M.S.; Lee, J.M.; Kim, T.Y.; Park, Y.S.; Ma, E. Mechanisms of Li+ transport in garnet-type cubic Li3+xLa3M2O12 (M. = Te, Nb, Zr). Phys. Rev. B 2012, 85, 052301. [Google Scholar] [CrossRef]
  18. Murugan, R.; Thangadurai, V.; Weppner, W. Fast Lithium ion conduction in garnet-type Li7La3Zr2O12. Angew. Chem. Int. Ed. 2007, 46, 7778. [Google Scholar] [CrossRef]
  19. Bernstein, N.; Johannes, M.D.; Hoang, K. Origin of the structural phase transition in Li7La3Zr2O12. Phys. Rev. Lett. 2012, 109, 205702. [Google Scholar] [CrossRef] [Green Version]
  20. Rangasamy, E.; Wolfenstine, J.; Sakamoto, J. The role of Al and Li concentration on the formation of cubic garnet solid electrolyte of nominal composition Li7La3Zr2O12. Solid State Ion. 2012, 206, 28. [Google Scholar] [CrossRef]
  21. Buannic, L.; Orayech, B.; López Del Amo, J.-M.; Carrasco, J.; Katcho, N.A.; Aguesse, F.; Manalastas, W.; Zhang, W.; Kilner, J.; Llordes, A. Dual substitution strategy to enhance Li+ ionic conductivity in Li7La3Zr2O12 solid electrolyte. Chem. Mater. 2017, 29, 1769. [Google Scholar] [CrossRef] [Green Version]
  22. Rettenwander, D.; Redhammer, G.; Preishuber-Pflugl, F.; Cheng, L.; Miyara, L.; Wagner, R.; Welzl, A.; Suard, E.; Doeff, M.M.; Wilkening, M.; et al. Structural and electrochemical consequences of Al and Ga cosubstitution in Li7La3Zr2O12 solid electrolytes. Chem. Mater. 2016, 28, 2384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Deviannapoorani, C.; Dhivya, L.; Ramakumar, S.; Murugan, R. Lithium ion transport properties of high conductive tellurium substituted Li7La3Zr2O12 cubic lithium garnets. J. Power Sources 2013, 240, 18. [Google Scholar] [CrossRef]
  24. Luo, Y.; Li, X.; Chen, H.; Guo, L. Influence of sintering aid on the microstructure and conductivity of the garnettype W-doped Li7La3Zr2O12 ceramic electrolyte. J. Mater. Sci. Mater. Electron. 2019, 30, 17195. [Google Scholar] [CrossRef]
  25. Wang, C.; Lin, P.P.; Gong, Y.; Liu, Z.G.; Lin, T.S.; He, P. Co-doping effects of Ba2+ and Ta5+ on the microstructure and ionic conductivity of garnet-type solid state electrolytes. J. Alloys Compd. 2021, 854, 157143. [Google Scholar] [CrossRef]
  26. Han, J.; Zhu, J.; Li, Y.; Yu, X.; Wang, S.; Wu, G.; Xie, H.; Vogel, S.C.; Izumi, F.; Momma, K.; et al. Experimental visualization of of lithium conduction pathways in garnet-type Li7La3Zr2O12. Chem. Commun. 2012, 48, 9840. [Google Scholar] [CrossRef]
  27. Wang, D.; Zhong, G.; Dolotko, O.; Li, Y.; McDonald, M.J.; Mi, J.; Fu, R.; Yang, Y. The synergistic effects of Al and Te on the structure and Li+-mobility of garnet-type solid electrolytes. J. Mater. Chem. A 2014, 2, 20271. [Google Scholar] [CrossRef]
  28. Izumi, F.; Ikeda, T. A Rietveld-Analysis Programm RIETAN-98 and Its Applications to Zeolites. Mater. Sci. Forum. 2000, 321–324, 198–205. [Google Scholar] [CrossRef]
  29. Almond, D.P.; Ducan, G.K.; West, A.R. The determination of hopping rates and carrier concentrations in ionic conductors by a new analysis of ac conductivity. Solid State Ion. 1983, 8, 159–164. [Google Scholar] [CrossRef]
  30. Ahmad, M.M.; Yamada, K. Hopping rates and concentration of mobile fluoride ions in Pb1-xSnxF2 solid solutions. J. Chem. Phys. 2007, 127, 124507. [Google Scholar] [CrossRef]
  31. Hairetdinov, E.F.; Uvarov, N.F.; Patel, H.K.; Martin, S.W. Estimation of the free-charge-carrier concentration in fast-ion conducting Na2S-B2S3 glasses from as analysis of the frequency-dependent conductivity. Phys. Rev. B 1994, 50, 13259–13266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Ghosh, A.; Pan, A. Scaling of the conductivity spectra in ionic glasses: Dependence on the structure. Phys. Rev. Lett. 2000, 84, 2188. [Google Scholar] [CrossRef] [PubMed]
  33. Adams, S.; Rao, R.P. Ion transport and phase transition in Li7-xLa3(Zr2-xMx)O12 (M = Ta5+, Nb5+, x = 0, 0.25). J. Mater. Chem. 2012, 22, 1426–1434. [Google Scholar] [CrossRef]
  34. Ramirez, A.P.; Subramanian, M.A.; Gardel, M.; Blumberg, G.; Li, D.; Vogt, T.; Shapiro, S.M. Giant dielectric constant response in a copper-titanate. Solid State Commun. 2000, 115, 217–220. [Google Scholar] [CrossRef]
  35. Li, M.; Sinclair, D.C.; West, A.R. Extrinsic origins of the apparent relaxorlike behavior in CaCu3Ti4O12 ceramics at high temperatures: A cautionary tale. J. Appl. Phys. 2011, 109, 084106. [Google Scholar] [CrossRef]
  36. Ahmad, M.M.; Kotb, H.M.; Joseph, C.; Kumar, S.; Alshoaibi, A. Transport and dielectric properties of mechanosynthesized La2/3Cu3Ti4O12 ceramics. Crystals 2021, 11, 313. [Google Scholar] [CrossRef]
  37. Garcia-Martin, S.; Morata-Orrantia, A.; Alario-Franco, M.A. Influence of the crystal microstructure on the dielectric response of the La0.67Li0.2Ti0.8Al0.2O3. J. Appl. Phys. 2006, 100, 054101. [Google Scholar] [CrossRef]
  38. Ahmad, M.M. Giant dielectric properties of nanocrystalline Pb1-xSnxF2 solid solutions. J. Mater. Sci. Mater. Electron. 2014, 25, 4398–4403. [Google Scholar] [CrossRef]
Figure 1. Rietveld refinement of XRD data of Li5+2xLa3Ta2-xGaxO12 garnets.
Figure 1. Rietveld refinement of XRD data of Li5+2xLa3Ta2-xGaxO12 garnets.
Crystals 12 00770 g001
Figure 2. SEM micrographs of Li5+2xLa3Ta2-xGaxO12 ceramics with different Ga+3 compositions.
Figure 2. SEM micrographs of Li5+2xLa3Ta2-xGaxO12 ceramics with different Ga+3 compositions.
Crystals 12 00770 g002
Figure 3. Complex impedance diagrams at selected temperatures for Li5.6La3Ta1.7Ga0.3O12 garnets at (a) low and (b) high temperature regions.
Figure 3. Complex impedance diagrams at selected temperatures for Li5.6La3Ta1.7Ga0.3O12 garnets at (a) low and (b) high temperature regions.
Crystals 12 00770 g003
Figure 4. The temperature dependence of the total ionic conductivity for different compositions of Li5+2xLa3Ta2-xGaxO12 garnets along with the pure Li5La3Ta2O12.
Figure 4. The temperature dependence of the total ionic conductivity for different compositions of Li5+2xLa3Ta2-xGaxO12 garnets along with the pure Li5La3Ta2O12.
Crystals 12 00770 g004
Figure 5. The conductivity spectra of LLT-Ga garnets with x = 0.30 at different temperatures. The solid curves between the points represent the best fits to Equation (2).
Figure 5. The conductivity spectra of LLT-Ga garnets with x = 0.30 at different temperatures. The solid curves between the points represent the best fits to Equation (2).
Crystals 12 00770 g005
Figure 6. (a) The temperature dependence of the dc conductivity, σdc, and (b) the hopping frequency, ωH, as determined from the fitting process in Figure 5.
Figure 6. (a) The temperature dependence of the dc conductivity, σdc, and (b) the hopping frequency, ωH, as determined from the fitting process in Figure 5.
Crystals 12 00770 g006
Figure 7. (a) ε′ spectra for LLT-Ga1 at selected temperatures as a representative sample and (b) the Cole-Cole diagram at 280 K for LLT-Ga1 with the inset showing high magnification of the high frequency region. (c) ε′ spectra for all Li5+2xLa3Ta2-xGaxO12 compositions at 300 K.
Figure 7. (a) ε′ spectra for LLT-Ga1 at selected temperatures as a representative sample and (b) the Cole-Cole diagram at 280 K for LLT-Ga1 with the inset showing high magnification of the high frequency region. (c) ε′ spectra for all Li5+2xLa3Ta2-xGaxO12 compositions at 300 K.
Crystals 12 00770 g007
Figure 8. (a) The frequency dependence of tan δ for LLT-Ga1 as a representative example, and (b) the temperature dependence of the dielectric relaxation time of Li5+2xLa3Ta2-xGaxO12 garnets.
Figure 8. (a) The frequency dependence of tan δ for LLT-Ga1 as a representative example, and (b) the temperature dependence of the dielectric relaxation time of Li5+2xLa3Ta2-xGaxO12 garnets.
Crystals 12 00770 g008
Table 1. The values of σdc at 300 K and the activation energy ΔE for ionic conduction of Li5+2xLa3Ta2-xGaxO12 garnets. Eσ and EH are the activation energy for the conduction and the hopping process, whereas Etanδ is the activation energy of the dielectric relaxation process determined from dissipation factor. nc is the average value of the concentration of Li+ mobile ions.
Table 1. The values of σdc at 300 K and the activation energy ΔE for ionic conduction of Li5+2xLa3Ta2-xGaxO12 garnets. Eσ and EH are the activation energy for the conduction and the hopping process, whereas Etanδ is the activation energy of the dielectric relaxation process determined from dissipation factor. nc is the average value of the concentration of Li+ mobile ions.
 
x
σdc at 300 K
(S/cm)
ΔE
(e.V)
Eσ
(e.V)
EH
(e.V)
Etanδ
(e.V)
nc
(cm−3)
0.004.26 × 10−60.620.620.550.519.10 × 1020
0.103.21 × 10−50.550.530.470.551.23 × 1021
0.203.13 × 10−50.550.520.460.541.31 × 1021
0.304.00 × 10−50.560.530.470.571.21 × 1021
0.403.68 × 10−50560.530.470.571.62 × 1021
0.501.50 × 10−50.560.550.520.569.92 × 1020
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmad, M.M.; Al-Ghareeb, F.R.; Kotb, H.M.; Ansari, S.A.; Kayed, T.S.; Khater, H.A.; Kumar, S.; Yamada, K. Enhanced Li+ Ionic Conduction and Relaxation Properties of Li5+2xLa3Ta2-xGaxO12 Garnets. Crystals 2022, 12, 770. https://doi.org/10.3390/cryst12060770

AMA Style

Ahmad MM, Al-Ghareeb FR, Kotb HM, Ansari SA, Kayed TS, Khater HA, Kumar S, Yamada K. Enhanced Li+ Ionic Conduction and Relaxation Properties of Li5+2xLa3Ta2-xGaxO12 Garnets. Crystals. 2022; 12(6):770. https://doi.org/10.3390/cryst12060770

Chicago/Turabian Style

Ahmad, Mohamad M., Fatimah R. Al-Ghareeb, Hicham Mahfoz Kotb, Sajid Ali Ansari, Tarek S. Kayed, Hassan A. Khater, Shalendra Kumar, and Koji Yamada. 2022. "Enhanced Li+ Ionic Conduction and Relaxation Properties of Li5+2xLa3Ta2-xGaxO12 Garnets" Crystals 12, no. 6: 770. https://doi.org/10.3390/cryst12060770

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