Next Article in Journal
Enhancing Optically Pumped Organic-Inorganic Hybrid Perovskite Amplified Spontaneous Emission via Compound Surface Plasmon Resonance
Next Article in Special Issue
Analysis of Waveguides on Lithium Niobate Thin Films
Previous Article in Journal
Fluorine Translational Anion Dynamics in Nanocrystalline Ceramics: SrF2-YF3 Solid Solutions
Previous Article in Special Issue
Structural and Magnetic Behavior of Oxidized and Reduced Fe Doped LiNbO3 Powders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computer Modelling of Hafnium Doping in Lithium Niobate

by
Romel M. Araujo
1,
Mario E. G. Valerio
2 and
Robert A. Jackson
3,*
1
Chemistry Coordination/IPISE/PIC, Pio Decimo College, Campus III, Aracaju-SE 49027-210, Brazil
2
Physics Department, Federal University of Sergipe, Campus Universitário, São Cristovão-SE 491000-000, Brazil
3
Lennard-Jones Laboratories, School of Chemical and Physical Sciences, Keele University, Keele, Staffordshire ST5 5BG, UK
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(3), 123; https://doi.org/10.3390/cryst8030123
Submission received: 3 January 2018 / Revised: 28 February 2018 / Accepted: 1 March 2018 / Published: 6 March 2018

Abstract

:
Lithium niobate (LiNbO3) is an important technological material with good electro-optic, acousto-optic, elasto-optic, piezoelectric and nonlinear properties. Doping LiNbO3 with hafnium (Hf) has been shown to improve the resistance of the material to optical damage. Computer modelling provides a useful means of determining the properties of doped and undoped LiNbO3, including its defect chemistry, and the effect of doping on the structure. In this paper, Hf-doped LiNbO3 has been modelled, and the final defect configurations are found to be consistent with experimental results.

1. Introduction

Lithium niobate (LiNbO3) is a material with many important technological applications that result from its diverse physical properties [1,2,3,4]. Laser-induced optical damage or so-called photorefraction was first observed in LiNbO3 and LiTiO3 crystals at the Bell Laboratories [5]. This effect can be utilized for holographic information storage and optical amplification; however, it hinders the usage of LiNbO3 in frequency doublers, Q-switchers and optical waveguides, so ways of minimising this optical damage have been sought actively. Kokanyan et al. [6] reported that the light-induced birefringence changes of LiNbO3 crystals doped with 4 mol % of HfO2 were comparable to that of 6 mol % MgO doped crystals, indicating that Hf doping is effective in resisting optical damage.
Much useful information about lithium niobate and its defect properties can be obtained by computer modelling, based on the description of interactions between ions by effective potentials. Previous papers have reported the derivation of an interatomic potential for LiNbO3 [7], the doping of the structure by rare earth ions [8,9], doping with Sc, Cr, Fe and In [10], and metal co-doping [11]. These papers show that modelling can predict the energetically optimal locations of the dopant ions, and calculate the energy involved in the doping process, making it a suitable method to study Hf-doped lithium niobate, with the aim of establishing the optimal doping site and charge compensation scheme.

2. Methodology

In this paper, use is made of the lattice energy minimisation method, in which the lattice energy of a given structure is calculated, and the structure varied until a minimum in the energy is found. This approach has been applied to a wide range of inorganic materials, with specific applications to LiNbO3 reported in references [7,8,9,10,11]. The method makes use of interatomic potentials to describe the interactions between ions in the solid, as described in the next Section 2.1. Defects in solids are modelled using the Mott-Littleton method [12] which is described in Section 2.2. All calculations were performed using the GULP code [13].

2.1. Interatomic Potentials

In this paper use has been made of a previously derived potential for LiNbO3 [7], and a potential fitted to the structure of HfO2. In both cases, a Buckingham potential is employed, supplemented by an electrostatic interaction term:
V = q 1 q 2 r + A exp ( r ρ ) C r 6
In this potential, q1 and q2 are the charges on the interacting ions separated by a distance r, and A, ρ and C are parameters that are fitted empirically.
The derivation of potentials for LiNbO3 and HfO2 are considered separately below.

2.1.1. LiNbO3

Full details of the derivation of the LiNbO3 potential are given in reference [7], but they will be summarised here. The potential was derived empirically by simultaneously fitting to the structures of LiNbO3, Li2O and Nb2O5. The O2−–O2− potential obtained by Catlow [14] was retained as this is widely used in many other oxides. The O2− ion was described using the shell model [15], and a 3 body potential was used to model the interactions between niobium ions and nearest oxygen neighbours, which takes the form:
V3 body = ½ k (θ − θ0)2
In this equation θ0 is the equilibrium bond angle and kθ is the bond-bending force constant.
The potential parameters are given in Table 1 below:
A comparison of experimental [16] and calculated lattice parameters of LiNbO3 can be found in Table 2, showing that the derived potential reproduces the structural parameters to within a few percent.

2.1.2. HfO2

A potential was derived for HfO2 by fitting to its structure [17]. The potential parameters are given in Table 3 (with the O2− shell parameters having the same values as in LiNbO3), and the agreement between calculated and experimental lattice parameters calculated at 0 K and 293 K is shown in Table 4. As is seen from the ∆% values, good agreement is obtained using this potential.

2.2. Defect Calculations

The calculations are carried out using the Mott–Littleton method [12], in which point defects are considered to be at the centre of a region in which all interactions are treated explicitly, while approximate methods are employed for regions of the lattice more distant from the defect. In practice, this involves placing the Hf4+ ion at either the Li+ or Nb5+ site, along with a range of charge compensating defects, as listed below, using schemes (i) and (ii) suggested by Li et al. [18], plus a further 5 schemes ((iii)–(vii)) proposed here:
(i)
An Hf4+ ion at a Li+ site, with charge compensation by 3 Li+ vacancies
(ii)
An Hf4+ ion at a Li+ site, with charge compensation by 3 Hf4+ ions at Nb5+ sites
(iii)
4 Hf4+ ions at Nb5+ sites, with charge compensation by a Nb5+ ion at a Li+ site
(iv)
An Hf4+ ion at a Nb5+ site, with charge compensation by a Nb5+ ion at a Li+ site and 3 Li+ vacancies
(v)
2 Hf4+ ions at Nb5+ sites, with charge compensation by a Nb5+ ion at a Li+ site and 2 Li+ vacancies
(vi)
3 Hf4+ ions at Nb5+ sites, with charge compensation by a Nb5+ ion at a Li+ site and 1 Li+ vacancy
(vii)
2 Hf4+ ions at Nb5+ sites, with charge compensation by an O2− vacancy

3. Results and Discussion

The seven mechanisms described in Section 2.2 have been written below as solid-state reactions, employing Kroger–Vink notation [19]:
H f O 2 + 4 L i L i H f L i + 3 V L i + 2 L i 2 O
4 H f O 2 + L i L i + 3 N b N b H f L i + 3 H f N b + 1 2 L i 2 O + 3 2 N b 2 O 5
4 H f O 2 + L i L i + 4 N b N b 4 H f N b + N b L i + 1 2 L i 2 O + 3 2 N b 2 O 5
H f O 2 + 4 L i L i + N b N b H f N b + N b L i + 3 V L i + 2 L i 2 O
2 H f O 2 + 3 L i L i + 2 N b N b 2 H f N b + N b L i + 2 V L i + L i 2 O + L i N b O 3
3 H f O 2 + 2 L i L i + 3 N b N b 3 H f N b + N b L i + V L i + 2 L i N b O 3
2 H f O 2 + 2 N b N b + O O 2 H f N b + V O + N b 2 O 5
The energies corresponding to these reactions are defined as solution energies, Es, and they are calculated as follows:
E s = E ( H f L i + 3 V L i ) + 2 E l a t t ( L i 2 O ) E l a t t ( H f O 2 )
E s = E ( H f L i + 3 H f N b ) + 3 2 E l a t t ( N b 2 O 5 ) + 1 2 E l a t t ( L i 2 O ) 4 E l a t t ( H f O 2 )
E s = E ( 4 H f N b + N b L i ) + 3 2 E l a t t ( N b 2 O 5 ) + 1 2 E l a t t ( L i 2 O ) 4 E l a t t ( H f O 2 )
E s = E ( H f N b + N b L i + 3 V L i ) + 2 E l a t t ( L i 2 O ) E l a t t ( H f O 2 )
E s = E ( 2 H f N b + N b L i + 2 V L i ) + E l a t t ( L i 2 O ) + E l a t t ( L i N b O 3 ) 2 E l a t t ( H f O 2 )
E s = E ( 3 H f N b + N b L i + V L i ) + 2 E l a t t ( L i N b O 3 ) 3 E l a t t ( H f O 2 )
E s = E ( 2 H f N b + V O ) + E l a t t ( N b 2 O 5 ) 2 E l a t t ( H f O 2 )
Lattice energies, Elatt, required to calculate the solution energies, are given in Table 5. Table 6 gives the formation energies of the bound defects (the first term in the above equations). Table 7 gives the solution energy for each scheme (determined using the expressions above), and it is noted that the lowest energy corresponds to scheme (vi), where 3 Hf4+ ions substitute at Nb5+ sites, with charge compensation by a Nb5+ ion at a Li+ site and 1 Li+ vacancy, and the second lowest energy scheme is (ii), which involves the Hf4+ ion substituting at both cation sites (self-compensation). Experimental data [18,20] supports the self-compensation model, and it is noted that at 293 K the calculated energetic preference for scheme (vi) is only 0.06 eV. However, it is noted that the calculations in this paper have been made at infinite dilution, and experimental data suggests that if dopant concentration is taken into account, Hf in low concentrations occupies the Li+ site [21,22,23], and that occupancy of both Li+ and Nb5+ sites only happens once the optical damage threshold is passed [18,23]. Future calculations will be carried out which will model the effect of Hf concentration on the preferred dopant sites, enabling comparison with these results to be made.

4. Conclusions

This paper has presented a computational study of Hf4+-doped LiNbO3. Solution energies have been calculated for seven possible mechanisms by which the Hf4+ might be incorporated in the structure, and the lowest energy scheme, involving self-compensation, is shown to be consistent with some experimental data, although future calculations including Hf concentration will be carried out to investigate this further.

Acknowledgments

The authors are grateful to CAPES and FINEP for financial support.

Author Contributions

Romel M. Araujo carried out the calculations, Mario E. G. Valerio and Robert A. Jackson analysed the results and Robert A. Jackson wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jazbinsek, M.; Zgonik, M. Material tensor parameters of LiNbO3 relevant for electro-and elasto-optics. Appl. Phys. B 2002, 74, 407. [Google Scholar]
  2. Rauber, A. Current Topics in Materials Science; Kaldis, E., Ed.; North-Holland: Amsterdam, The Netherlands, 1978; Volume 1, p. 481. [Google Scholar]
  3. Prokhorov, A.M.; Kuzminov, Y.S. Physics and Chemistry of Crystalline Lithium Niobate, 1st ed.; Hilger: New York, NY, USA, 1990. [Google Scholar]
  4. Wong, K.K. (Ed.) Properties of Lithium Niobate; IEE: London, UK, 2004. [Google Scholar]
  5. Askin, A.; Boyd, G.D.; Dziedzic, J.M.; Smith, R.G.; Ballman, A.A.; Levinstein, J.J.; Nassau, K. Optically-induced refractive index inhomogeneities in LiNbO3, and LiTaO3. Appl. Phys. Lett. 1966, 9, 72. [Google Scholar] [CrossRef]
  6. Kokanyan, E.P.; Razzari, L.; Cristiani, I.; Degiorgio, V.; Gruber, J.B. Reduced photorefraction in hafnium-doped single-domain and periodically poled lithium niobate crystals. Appl. Phys. Lett. 2004, 84, 1880. [Google Scholar] [CrossRef]
  7. Jackson, R.A.; Valerio, M.E.G. A new interatomic potential for the ferroelectric and paraelectric phases of LiNbO3. J. Phys. Condens. Matter 2005, 17, 837. [Google Scholar] [CrossRef]
  8. Araujo, R.M.; Lengyel, K.; Jackson, R.A.; Valerio, M.E.G.; Kovacs, L. Computer modelling of intrinsic and substitutional defects in LiNbO3. Phys. Status Solidi (c) 2007, 4, 1201–1204. [Google Scholar] [CrossRef]
  9. Araujo, R.M.; Lengyel, K.; Jackson, R.A.; Kovacs, L.; Valerio, M.E.G. A computational study of intrinsic and extrinsic defects in LiNbO3. J. Phys. Condens. Matter 2007, 19, 046211. [Google Scholar] [CrossRef]
  10. Araujo, R.M.; Valerio, M.E.G.; Jackson, R.A. Computer modelling of trivalent metal dopants in lithium niobite. J. Phys. Condens. Matter 2008, 20, 035201. [Google Scholar] [CrossRef]
  11. Araujo, R.M.; Valerio, M.E.G.; Jackson, R.A. Computer modelling of hafnium doping in lithium niobite. Proc. R. Soc. A 2014, 470, 0406. [Google Scholar] [CrossRef]
  12. Mott, N.F.; Littleton, M.J. Conduction in polar crystals. I. Electrolytic conduction in solid salts. Trans. Faraday Soc. 1938, 34, 485–499. [Google Scholar] [CrossRef]
  13. Gale, J.D. GULP: A computer program for the symmetry-adapted simulation of solids. J. Chem. Soc. Faraday Trans. 1997, 93, 629–637. [Google Scholar] [CrossRef]
  14. Sanders, M.J.; Leslie, M.; Catlow, C.R.A. Interatomic potentials for SiO2. J. Chem. Soc. Chem. Commun. 1984, 1271–1273. [Google Scholar] [CrossRef]
  15. Dick, B.J.; Overhauser, A.W. Theory of the dielectric constants of alkali halide crystals. Phys. Rev. 1958, 112, 90. [Google Scholar] [CrossRef]
  16. Abrahams, S.C.; Marsh, P. Defect structure dependence on composition in lithium niobite. Acta Crystallogr. B 1986, 42, 61–68. [Google Scholar] [CrossRef]
  17. Liu, Q.J.; Liu, Z.T.; Feng, L.P.; Xu, B. Electronic structure, effective masses, mechanical and thermo-acoustic properties of cubic HfO2 under pressure. Phys. Status Solidi B 2011, 248, 950–955. [Google Scholar] [CrossRef]
  18. Li, S.; Liu, S.; Kong, Y.; Deng, D.; Gao, G.; Li, Y.; Gao, H.; Zhang, L.; Hang, Z.; Chen, S.; et al. The optical damage resistance and absorption spectra of LiNbO3: Hf crystals. J. Phys. Condens. Matter 2006, 18, 3527–3534. [Google Scholar] [CrossRef]
  19. Kröger, F.A.; Vink, H.J. The origin of the fluorescence in self-activated ZnS, CdS, and ZnO. J. Chem. Phys. 1954, 22, 250. [Google Scholar] [CrossRef]
  20. Marques, J.G.; Kling, A.; de Jesus, C.M.; Soares, J.C.; da Silva, M.F.; Dieguez, E.; Agulló-Lopez, F. Annealing recovery of neutron irradiated LiNbO3: Hf single crystals. Nucl. Instrum. Methods Phys. Res. B 1998, 141, 326–331. [Google Scholar] [CrossRef]
  21. Rebouta, L.; Soares, J.C.; Da Silva, M.F.; Sanz Garcia, J.A.; Dieguez, E.; Agulló-Lopez, F. Determination of lattice sites for Eu, Hf and Nd IN LiNbO3 by RBS/channeling experiments. Nucl. Instrum. Methods Phys. Res. B 1990, 50, 428–430. [Google Scholar] [CrossRef]
  22. Prieto, C.; Zaldo, C.; Fessler, P.; D’Expert, H.; Sanz Garcia, J.A.; Dieguez, E. Lattice position of Hf and Ta in LiNbO3: An extended X-ray-absorption fine-structure study. Phys. Rev. B 1991, 43, 2594–2600. [Google Scholar] [CrossRef]
  23. Hammoum, R.; Fontana, M.D.; Gilliot, M.; Bourson, P.; Kokanyan, E.P. Site spectroscopy of Hf doping in Hf-doped LiNbO3 crystals. Solid State Commun. 2009, 149, 1967–1970. [Google Scholar] [CrossRef]
Table 1. Potential parameters for LiNbO3 [7].
Table 1. Potential parameters for LiNbO3 [7].
InteractionA (eV)ρ (Å)C (eV Å6)
Nbcore–Oshell1425.00.36500.0
Licore–Oshell950.00.26100.0
Oshell–Oshell22,764.00.149027.88
Shell ParametersShell Charge, Y (|e|)Spring Constant, kr (eV Å−2)
O2−−2.970.0
3 body ParametersForce Constant, kθ (eV rad−2)Equilibrium Angle, θ0
Oshell–Nbcore–Oshell0.577690.0
Table 2. Comparison of experimental [16] and calculated lattice parameters for LiNbO3.
Table 2. Comparison of experimental [16] and calculated lattice parameters for LiNbO3.
ParameterExperimentalCalculated (0 K)∆%Calculated (295 K)∆%
a = b (Å)5.14745.15590.175.18680.77
c (Å)13.856113.68341.2413.71031.05
Table 3. Interionic potentials obtained from a fit to the HfO2 structure [17].
Table 3. Interionic potentials obtained from a fit to the HfO2 structure [17].
InteractionA (eV)ρ (Å)C (eV Å6)
Hfcore–Oshell1413.540.35090.0
Oshell–Oshell22764.00.149027.88
Table 4. Comparison of calculated and experimental lattice parameters.
Table 4. Comparison of calculated and experimental lattice parameters.
ParameterExperimental [17]Calculated (0 K)∆%Calculated (295 K)∆%
a = b = c (Å)5.0840005.0842360.005.0871190.06
Table 5. Lattice energies used in the solution energy calculations (eV).
Table 5. Lattice energies used in the solution energy calculations (eV).
Structures0 K293 K
LiNbO3−174.57−174.66
Li2O−33.16−32.92
Nb2O5−314.37−313.99
HfO2−110.39−110.45
Table 6. Defect formation energies, in eV, for the bound defect.
Table 6. Defect formation energies, in eV, for the bound defect.
DefectScheme (i)Scheme (ii)Scheme (iii)Scheme (iv)Scheme (v)Scheme (vi)Scheme (vii)
T (K)0293029302930293029302930293
Hf4+−36.03−36.3552.5152.2753.3453.11−33.85−34.01−2.61−2.7325.4025.0797.8297.64
Table 7. Solution energies, in eV, for the bound defect (per dopant ion).
Table 7. Solution energies, in eV, for the bound defect (per dopant ion).
DefectScheme (i)Scheme (ii)Scheme (iii)Scheme (iv)Scheme (v)Scheme (vi)Scheme (vii)
T (K)0293029302930293029302930293
Hf4+8.268.041.651.481.861.695.255.112.122.091.491.422.282.11

Share and Cite

MDPI and ACS Style

Araujo, R.M.; Valerio, M.E.G.; Jackson, R.A. Computer Modelling of Hafnium Doping in Lithium Niobate. Crystals 2018, 8, 123. https://doi.org/10.3390/cryst8030123

AMA Style

Araujo RM, Valerio MEG, Jackson RA. Computer Modelling of Hafnium Doping in Lithium Niobate. Crystals. 2018; 8(3):123. https://doi.org/10.3390/cryst8030123

Chicago/Turabian Style

Araujo, Romel M., Mario E. G. Valerio, and Robert A. Jackson. 2018. "Computer Modelling of Hafnium Doping in Lithium Niobate" Crystals 8, no. 3: 123. https://doi.org/10.3390/cryst8030123

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