research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Volume 48| Part 6| December 2015| Pages 1896-1900

Atomic disorder of Li0.5Ni0.5O thin films caused by Li doping: estimation from X-ray Debye–Waller factors

CROSSMARK_Color_square_no_text.svg

aSynchrotron X-ray Station at SPring-8, National Institute for Materials Science (NIMS),1-1-1 Kouto, Sayo, Hyogo 679-5148, Japan, bSynchrotron X-ray Group, Quantum Beam Unit, NIMS, 1-1-1 Kouto, Sayo, Hyogo 679-5148, Japan, cDepartment of Innovative and Engineered Materials, Tokyo Institute of Technology, 4259-J3-16 Nagatsuta-cho, Midori-ku, Yokohama, Kanagawa 226-8502, Japan, and dGlobal Research Center for Environment and Energy Based Nanomaterials Science, Lithium Air Battery Specially Promoted Research Team, NIMS, 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan
*Correspondence e-mail: alyang@semi.ac.cn, SAKATA.Osami@nims.go.jp

Edited by G. Renaud, CEA-Grenoble DSM/INAC/SP2M/NRS, Grenoble, France (Received 7 April 2015; accepted 22 October 2015; online 19 November 2015)

Cubic type room-temperature (RT) epitaxial Li0.5Ni0.5O and NiO thin films with [111] orientation grown on ultra-smooth sapphire (0001) substrates were examined using synchrotron-based thin-film X-ray diffraction. The 1[\overline{1}]1 and 2[\overline{2}]2 rocking curves including six respective equivalent reflections of the Li0.5Ni0.5O and NiO thin films were recorded. The RT B1 factor, which appears in the Debye–Waller factor, of a cubic Li0.5Ni0.5O thin film was estimated to be 1.8 (4) Å2 from its 1[\overline{1}]1 and 2[\overline{2}]2 reflections, even though the Debye model was originally derived on the basis of one cubic element. The corresponding Debye temperature is 281 (39) K. Furthermore, the B2 factor in the pseudo-Debye–Waller factor is proposed. This parameter, which is evaluated using one reflection, was also determined for the Li0.5Ni0.5O thin film by treating Li0.5Ni0.5O and NiO as ideal NaCl crystal structures. A structural parameter for the atomic disorder is introduced and evaluated. This parameter includes the combined effects of thermal vibration, interstitial atoms and defects caused by Li doping using the two Debye–Waller factors.

1. Introduction

LixNi1-xO has attracted considerable attention recently as a potential material for high-performance electrochromic devices (Moulki et al., 2012[Moulki, H., Park, D. H., Min, B.-K., Kwon, H., Hwang, S.-J., Choy, J.-H., Toupance, T., Campet, G. & Rougier, A. (2012). Electrochim. Acta, 74, 46-52.]), UV detectors (Ohta et al., 2003[Ohta, H., Hirano, M., Nakahara, K., Maruta, H., Tanabe, T., Kamiya, M., Kamiya, T. & Hosono, H. (2003). Appl. Phys. Lett. 83, 1029-1031.]) and gas sensors (Garduno-Wilches & Alonso, 2013[Garduno-Wilches, I. & Alonso, J. C. (2013). Int. J. Hydrogen Energy, 38, 4213-4219.]). Recently, cubic type NiO thin films with (111) orientation containing large amounts of Li (up to 50 mol%) were epitaxically grown on ultra-smooth sapphire (0001) substrates by room-temperature (RT) pulsed laser deposition (Shiraishi et al., 2010[Shiraishi, N., Kato, Y., Arai, H., Tsuchimine, N., Kobayashi, S., Mitsuhashi, M., Soga, M., Kaneko, S. & Yoshimoto, M. (2010). Jpn. J. Appl. Phys. 49, 108001.]; Yang et al., 2014[Yang, A. L., Sakata, O., Yamauchi, R., Katsuya, Y., Kumara, L. S. R., Shimada, Y., Matsuda, A. & Yoshimoto, M. (2014). Appl. Surf. Sci. 320, 787-790.]). The ultra-smooth sapphire substrate used in this work has an atomic `step and terrace' structure on its surface, which consists of atomically flat terraces comprising single oxygen layers separated by periodic atomic steps with 0.22 nm height corresponding to the oxygen layer spacing along the c axis. These films could be grown even though bulk NiO containing over 30 mol% Li undergoes a phase transformation from the cubic to rhombohedral structure (Goodenough et al., 1958[Goodenough, J. B., Wickham, D. G. & Croft, W. J. (1958). J. Phys. Chem. Solids, 5, 107-116.]).

The Debye temperature is an important characteristic structural parameter of a solid and describes the dynamic motions of atoms in the material. A number of physical parameters such as mean-square atomic displacement (Herbstein, 1961[Herbstein, F. H. (1961). Adv. Phys. 10, 313-355.]) and elastic constant (Gazzara & Middleton, 1964[Gazzara, C. P. & Middleton, R. M. (1964). Adv. X-ray Anal. 7, 14-16.]) depend on the Debye temperature of a solid. Therefore, it is very important to determine the Debye temperature of LixNi1-xO in order to understand its physical properties, such as the change of heat capacity caused by Li doping, and develop related electronic devices. X-ray diffraction is a valuable method to study thin films and other epitaxial layer materials. The changes of the lattice constant and crystal quality caused by doping are frequently evaluated using this method (Dutta et al., 2010[Dutta, T., Gupta, P., Gupta, A. & Narayan, J. (2010). J. Appl. Phys. 108, 083715.]; Palatnikov et al., 2006[Palatnikov, M. N., Biryukova, I. V., Sidorov, N. V., Denisov, A. V., Kalinnikov, V. T., Smith, P. G. R. & Shur, V. Ya. (2006). J. Cryst. Growth, 291, 390-397.]). However, the effect of atomic disorder caused by doping elements on the X-ray diffraction intensity has seldom been estimated (Sakata & Hashizume, 1997[Sakata, O. & Hashizume, H. (1997). Acta Cryst. A53, 781-788.]). In this study, we attempted to quantitatively analyze the effect of atomic disorder caused by Li doping on the X-ray diffraction intensity of an Li0.5Ni0.5O thin film.

2. History of Debye temperature estimation by X-ray diffraction

Originally, the Debye temperature was determined using X-ray diffraction and was strictly limited to monoatomic cubic crystals. The integrated intensity, I, from a cubic sample can be expressed as (James, 1962[James, R. W. (1962). The Optical Principles of the Diffraction of X-rays. Woodbridge: Ox Bow Press.]) I = [KL^{hkl}_{\rm p}\vert F \vert^2 D], where K is a constant, which is a normalized factor depending on the experimental arrangement, Lhklp is the Lorentz–polarization factor [(1+ \cos^22\theta_{hkl})/(2\sin^2\theta_{hkl} \cos\theta_{hkl})], h, k and l are the Miller indices of the diffraction plane, θ is the Bragg diffraction angle, [\vert F \vert] is the modulus of the structure factor, and D is the Debye–Waller factor, expressed as [\exp(-2B\sin^2\theta/\lambda^{2})]. Here (James, 1962[James, R. W. (1962). The Optical Principles of the Diffraction of X-rays. Woodbridge: Ox Bow Press.])

[{B} = 8{\pi}^2\overline{u^2} = {{24\hbar^2\pi^2T}\over{mk_{\rm B}{\theta}^2_{\rm M}}}{\left[\Phi(x)+{{x}\over{4}}\right]}\eqno(1)]

and

[\Phi(x) = {{1}\over{x}}\int\limits_0^x{{{y}\over{\exp(y)-1}}}\,{\rm d}y,\quad x = {\theta_{\rm M}}/{T}. \eqno(2)]

[ \overline{u^2}] is the mean-square displacement of atoms from the sites of the average lattice of the solid solution. m is the mass of the atom, kB is the Boltzmann constant, [\hbar] is Planck's constant divided by 2π, [\theta _{\rm M}] is the Debye temperature and T is the experimental temperature.

Later, the Debye theory was extended to apply to binary alloys, such as KCl and NaCl, whose specific heats can be fairly well described by the Debye–Waller formula with an appropriate value of the characteristic temperature (James, 1962[James, R. W. (1962). The Optical Principles of the Diffraction of X-rays. Woodbridge: Ox Bow Press.]). The structure factor for binary alloys is formally written as (Murakami, 1953[Murakami, T. (1953). J. Phys. Soc. Japan, 8, 36-40.])

[\eqalignno{F(h& k l) = r_{A}f_{A}\exp[{2i\pi(hx_{A}+ky_{A}+lz_{A})}] \exp(-B_{A}\sin^2\theta/\lambda^{2})\cr & + r_{B}f_{B}\exp[2i\pi(hx_{B}+ky_{B}+lz_{B})] \exp(-B_{B}\sin^2\theta/\lambda^2) .&(3)}]

Here f is the atomic scattering factor, x, y, z are the atomic positions in the unit cell, and the subscripts A and B refer to the two components of the alloy. rA and rB are the atomic fractions of A and B, respectively. To perform a simple approximation (Kulkarni & Bichile, 1977[Kulkarni, R. G. & Bichile, G. K. (1977). J. Appl. Phys. 48, 1844-1847.]; Wathore & Kulkarni, 1980[Wathore, T. N. & Kulkarni, R. G. (1980). J. Mater. Sci. 15, 2927-2930.]; Pathak & Trivedi, 1973[Pathak, P. D. & Trivedi, J. M. (1973). Acta Cryst. A29, 45-49.]), we let BA = BB = [\overline{B}], and in equation (1)[link] use [\overline{m}] = rAmA+rB mB. This approximation allows a quasi-Debye temperature to be obtained. The experimental results summarized by Lonsdale (1948[Lonsdale, K. (1948). Acta Cryst. 1, 142-149.]) are in reasonable agreement with this approximation.

3. Introduction of a structure parameter

The Debye temperature has been evaluated from X-ray intensities not only at different temperatures for a Bragg peak (Herbstein, 1961[Herbstein, F. H. (1961). Adv. Phys. 10, 313-355.]) but also from different reflections at a given temperature (Herbstein, 1961[Herbstein, F. H. (1961). Adv. Phys. 10, 313-355.]; Kulkarni & Bichile, 1977[Kulkarni, R. G. & Bichile, G. K. (1977). J. Appl. Phys. 48, 1844-1847.]). Here, a real Debye–Waller factor D1 for an Li0.5Ni0.5O thin film was obtained by X-ray diffraction using two different reflections, 1[\overline{1}]1 and 2[\overline{2}]2, with the assumption that the rock-salt lattice is a simple cubic structure composed of one kind of atom and the Li atoms occupy Ni substitutional positions without any atomic disorder. We also assume that the thermal disorder contributes to the ratio of the diffracted intensities between the 1[\overline{1}]1 and 2[\overline{2}]2 reflections for simplicity. Second, a pseudo-Debye–Waller factor D2 for the Li0.5Ni0.5O thin film was obtained from the X-ray 1[\overline{1}]1 intensity ratio between Li0.5Ni0.5O and NiO. Finally, we introduce an atomic disorder parameter exp(δ) (δ [ \leq] 0), including the combined effects of thermal vibration, interstitial atoms and defects expressed using D1, D2 and the reported NiO Debye temperature (Freer, 1981[Freer, R. (1981). J. Mater. Sci. 16, 3225-3227.]). This term is required because the positions of atoms are changed by Li doping. This process is outlined in Fig. 1[link].

[Figure 1]
Figure 1
The definition of a structure parameter [exp(δ)] to describe the effect of atomic disorder caused by Li doping on the X-ray diffraction intensity of an Li0.5Ni0.5O thin film.

4. Samples and experiment

LixNi1-xO thin films were epitaxically grown on ultra-smooth sapphire (0001) substrates (Yoshimoto et al., 1995[Yoshimoto, M., Maeda, T., Ohnishi, T., Koinuma, H., Ishiyama, O., Shinohara, M., Kubo, M., Miura, R. & Miyamoto, A. (1995). Appl. Phys. Lett. 67, 2615-2617.]) by RT pulsed laser deposition. Details of the deposition conditions have been reported elsewhere (Shiraishi et al., 2010[Shiraishi, N., Kato, Y., Arai, H., Tsuchimine, N., Kobayashi, S., Mitsuhashi, M., Soga, M., Kaneko, S. & Yoshimoto, M. (2010). Jpn. J. Appl. Phys. 49, 108001.]; Yang et al., 2014[Yang, A. L., Sakata, O., Yamauchi, R., Katsuya, Y., Kumara, L. S. R., Shimada, Y., Matsuda, A. & Yoshimoto, M. (2014). Appl. Surf. Sci. 320, 787-790.]). The compositions of NiO thin films containing large amounts of Li grown at RT were determined using inductively coupled plasma atomic emission spectrometry (Shimadzu ICPS-8100). The lithium contents x in the LixNi1-xO samples were 0 and 0.5. The periodicities of the intensity oscillations appearing in the X-ray reflectivity curves revealed that the thicknesses t of the NiO and Li0.5Ni0.5O epitaxial thin films were 400 (21) and 231 (6) Å, respectively.

X-ray diffraction measurements were performed with a six-axis diffractometer at the National Institute for Materials Science (NIMS) beamline BL15XU, SPring-8. The wavelength used was 1.000 Å.

5. Results and discussion

5.1. Crystal structure

The crystallographic epitaxy between the LixNi1-xO thin films and sapphire substrate was Li0.5Ni0.5O [1[\overline{2}]1] // sapphire [11[\overline{2}]0] and Li0.5Ni0.5O [111] // sapphire [0001]. The cube-on-hexagonal epitaxial relationship was the same as that reported previously for similar samples (Sakata et al., 2004[Sakata, O., Yi, M. S. Matsuda, A., Liu, J., Sato, S., Akiba, S., Sasaki, A. & Yoshimoto, M. (2004). Appl. Surf. Sci. 221, 450-454.]; Yang et al., 2014[Yang, A. L., Sakata, O., Yamauchi, R., Katsuya, Y., Kumara, L. S. R., Shimada, Y., Matsuda, A. & Yoshimoto, M. (2014). Appl. Surf. Sci. 320, 787-790.]). The φ scan around the nonspecular NiO (1[\overline{1}]1) Bragg positions showed sixfold symmetry (not shown here). This is because two types of terraces formed alternately on the sapphire (0001) face, with 0.2 nm-high steps (Yamauchi et al., 2012[Yamauchi, R., Tan, G., Shiojiri, D., Koyama, K., Kaneko, S., Matsuda, A. & Yoshimoto, M. (2012). Jpn. J. Appl. Phys. 51, 06FF16.]). We also recorded the X-ray intensities of six 1[\overline{1}]1 and 2[\overline{2}]2 diffraction peaks.

5.2. Debye temperature of an Li0.5Ni0.5O epitaxial thin film

Fig. 2[link] shows one of the six X-ray rocking curves around the 1[\overline{1}]1 Bragg peaks of the NiO and Li0.5Ni0.5O thin films. The average full width at half-maximum of the Li0.5Ni0.5O thin film (5.48°) was much larger than that of the NiO thin film (1.60°). This suggests that the crystal quality was lowered after Li doping.

[Figure 2]
Figure 2
X-ray rocking curves around the 1[\overline{1}]1 Bragg peaks of NiO and Li0.5Ni0.5O thin films.

The average integrated intensity ratio R1 between the 1[\overline{1}]1 and 2[\overline{2}]2 Bragg peaks of the Li0.5Ni0.5O epitaxial thin film was calculated to be 2.5 (2). The contribution of thermal diffuse scattering (TDS) to the measured intensity was neglected in our analysis, because the TDS correction was small (less than 4% even in imperfect single-crystal silicon; Matsumuro et al., 1990[Matsumuro, A., Kobayashi, M., Kikegawa, T. & Senoo, M. (1990). J. Appl. Phys. 68, 2719-2722.]; Chipman & Batterman, 1963[Chipman, D. R. & Batterman, B. W. (1963). J. Appl. Phys. 34, 912-913.]) in our samples compared with the statistical fluctuation, which was about 7.6% in the NiO thin film. If we assume that all Li atoms are substituted at Ni atom positions, the composition of the Li0.5Ni0.5O epitaxial thin film is uniform and structural disorder is neglected, the diffracted intensity from the Li0.5Ni0.5O epitaxial film is

[\eqalignno{I(hkl)^{\rm obs}_{\rm Li_{0.5}Ni_{0.5}O} \propto&\, |F(hkl)^{\rm cal}_{\rm Li_{0.5}Ni_{0.5}O} |^2\,L^{hkl}_{\rm p} \cr & \times t^2_{\rm Li_{0.5}Ni_{0.5}O}\exp[{-2B} (\sin\theta_{hkl}/\lambda)^2].&(4)}]

Let us introduce the ratio S1cal of the square of the calculated structure factors:

[S^{\rm cal}_1\equiv{{|F(1\overline{1}1)^{\rm cal}_{\rm Li_{0.5}Ni_{0.5}O }|^2}\over{|F(2\overline{2}2)^{\rm cal}_{\rm Li_{0.5}Ni_{0.5}O} |^2}}. \eqno(5)]

The values of Scal1 are estimated to be 0.331 and 0.328 when we assume that the NiO and Li0.5Ni0.5O films have the NaCl crystal structure and use the atomic scattering factors (Brown et al., 2006[Brown, P. J., Fox, A. G., Maslen, E. N., O'Keefe, M. A. & Willis, B. T. M. (2006). International Tables for Crystallography, Vol. C, Mathematical, Physical and Chemical Tables, 1st online ed., edited by E. Prince, ch. 6.1, pp. 554-595. Chester: International Union of Crystallography.]) for their neutral atoms (Li, Ni and O) and ions (Li+, Ni2+ and O2-), respectively. The calculated results were almost the same. Let us define the observed integrated intensity ratio R1:

[{R_1}\equiv{{I(1\overline{1}1)^{\rm obs}_{\rm Li_{0.5}Ni_{0.5}O }}\over{I(2\overline{2}2)^{\rm obs}_{\rm Li_{0.5}Ni_{0.5}O }}}.\eqno(6)]

The averaged value of R1 for the six equivalent reflections was 2.5 (2). Substituting equation (4)[link] into equation (6)[link] allows us to derive the formula for B1 as follows:

[\eqalignno{ B_1 = &\, - 0.5\ln (R_1 L^{2\overline{2}2}_{\rm p}/ S^{\rm cal}_1 L^{1\overline{1}1}_{\rm p})\cr & /[(\sin\theta_{1\overline{1}1}/\lambda)^2- (\sin\theta_{2\overline{2}2}/\lambda)^2].&(7)}]

The B1 factor of Li0.5Ni0.5O was calculated to be 1.8 (4) Å2 when we used parameters observed from the Li0.5Ni0.5O films: [\theta_{1\overline{1}1}] = 12.19° and [\theta_{2\overline{2}2}] = 24.88°, and (1[\overline{1}]1) lattice spacing = 2.368 (4) Å. The corresponding Debye temperature was 281 (39) K according to equations (1)[link] and (2)[link]. The error originates from that of B1, and it is a little larger than that already determined by X-ray diffraction (Herbstein, 1961[Herbstein, F. H. (1961). Adv. Phys. 10, 313-355.]), so it may be caused by the large statistical fluctuation of integrated intensity. The corresponding r.m.s. amplitude of atomic vibration was evaluated to be 0.15 Å, as shown in Table 1[link]. The ratio r between the r.m.s. amplitude of atomic vibration and the (1[\overline{1}]1) lattice spacing of the Li0.5Ni0.5O thin film was 6.3%. This value is larger than that of ideal bulk NiO (4.6%; Freer, 1981[Freer, R. (1981). J. Mater. Sci. 16, 3225-3227.]), which suggests that Li doping increased the lattice thermal vibration (Borca et al., 2000[Borca, C. N., Komesu, T., Jeong, H. K., Dowben, P. A., Ristoiu, D., Hordequin, C., Pierre, J. & Nozières, J. P. (2000). Appl. Phys. Lett. 77, 88-90.]). D1 is written as [\exp[-2B_1(\sin\theta_{1-11}/\lambda)^2]] and is here called the real Debye–Waller factor for the Li0.5Ni0.50 thin film.

Table 1
B1 and BNiO factors, Debye temperature θM, r.m.s. amplitude of atomic vibration [\overline {u^2}^{1/2}], ([1 \overline 1 1]) lattice spacing [d_{1 \overline 1 1}], and ratio [r = \overline {u^2}^{1/2} / d_{1 \overline 1 1}] of Li0.5Ni0.5O and NiO thin films at 298 K

Sample B1 or BNiO2) θM (K) [\overline{u^2}^{1/2}] (Å) [d_{1\overline{1}1}] (Å) r (%)
Li0.5Ni0.5O 1.8 (4) 281 (39) 0.15 (2) 2.368 (4) 6.3 (8)
NiO 0.939 317.4 (Freer, 1981[Freer, R. (1981). J. Mater. Sci. 16, 3225-3227.]) 0.11 2.395 (2) 4.6

5.3. Estimation of the effect of atomic disorder caused by Li doping on the X-ray diffraction intensity

In order to obtain the structure parameter that can be used to estimate the effect of atomic disorder [exp(δ)] caused by Li doping on the X-ray diffraction intensity of an Li0.5Ni0.5O thin film, let us propose a factor D2 from the X-ray 1[\overline{1}]1 intensity ratio between Li0.5Ni0.5O and NiO as shown in Fig. 1[link]. If we assume that all of the Li atoms are substituted for Ni atoms and the composition of the Li0.5Ni0.5O epitaxial thin film is uniform (Fig. 3[link]a), then structural disorder can be neglected.

[Figure 3]
Figure 3
Crystal structure models of Li0.5Ni0.5O with (a) all of the Li atoms substituted for Ni atoms and (b) all of the Li atoms located in interstitial sites.

Let us define the measured average intensity ratio R2 between the 1[\overline{1}]1 Bragg peaks of the Li0.5Ni0.5O and NiO films:

[R_2\equiv{{I(1\overline{1}1)^{\rm obs}_{\rm Li_{0.5}Ni_{0.5}O }}\over {I(1\overline{1}1)^{\rm obs}_{\rm NiO}}} .\eqno(8)]

The average value of R2 for the six equivalent reflections was 0.029 (4). We also define

[S^{\rm cal}_2\equiv{{|F(1\overline{1}1)^{\rm cal}_{\rm Li_{0.5}Ni_{0.5}O }|^2}\over{|F(1\overline{1}1)^{\rm cal}_{\rm NiO}|^2}}.\eqno(9)]

The value of Scal2 is 0.148 when we assume that the NiO and Li0.5Ni0.5O films have the NaCl crystal structure and use the atomic scattering factors for their neutral atoms. The formula for our proposed parameter B2 is obtained after we substitute equation (4)[link] into equation (8)[link]:

[\eqalignno{ B_2 = & - 0.5[\ln (R_2 L^{\rm NiO1\overline{1}1}_{\rm p} t^2_{\rm NiO}/ S^{\rm cal}_2 L^{1\overline{1}1}_{\rm p} t^2_{\rm Li_{0.5}Ni_{0.5}O}) \cr &-2B_{\rm NiO}(\sin\theta^{\rm NiO}_{1\overline{1}1}/\lambda)^2] /(\sin\theta_{1\overline{1}1}/\lambda)^2.&(10)}]

B2 was evaluated to be 6.5 (15) Å2 using the BNiO value of 0.939 Å2 obtained from [\theta _{\rm M}] (317.4 K at RT; Freer, 1981[Freer, R. (1981). J. Mater. Sci. 16, 3225-3227.]) of NiO according to equations (1)[link] and (2)[link]. We also used parameters ([\theta_{1\overline{1}1}^{\rm NiO}] = 12.05°) observed from the NiO film. These obtained parameters, the r.m.s. amplitude of atomic vibration and its ratio compared with the lattice spacing are listed in Table 2[link]. D2 is written as [\exp[-2B_2(\sin\theta_{1\overline 11}/\lambda)^2]] and is here called the pseudo-Debye–Waller factor for the Li0.5Ni0.50 thin film. B2 is much greater than B1. We do not consider the atomic disorder when obtaining the B1 value. To consider the difference between B1 and B2, let us assume the relation of D2 and D1 using the atomic disorder parameter expressed as [D_2 = \exp(\delta) D_1]. [\exp(\delta)] was evaluated to be 0.66 using NiO as reference. This value is much smaller than unity, which suggests that the range in which the atoms are located might be ca 0.14 Å wider than that for B1. In general, Debye–Waller factors may include not only the influence of the thermal disorder but also that of the atomic disorder. The structural disorder that we evaluated looks to have a very small influence on B1. Furthermore, let us assume another extreme case where all of the Li atoms are located on interstitial sites. Interstitial Li atoms are found not only in cubic Li-doped NiO (Guo et al., 2013[Guo, W., Hui, K. N. & Hui, K. S. (2013). Mater. Lett. 92, 291-295.]) but also in many other crystal structures like ZnO (Chawla et al., 2009[Chawla, S., Jayanthi, K. & Kotnala, R. K. (2009). Phys. Rev. B, 79, 125204.]), ZnSe (Sasaki et al., 1993[Sasaki, T., Oguchi, T. & Katayama-Yoshida, H. (1993). Physica B, 185, 159-163.]) and Li3NbO4 (McLaren et al., 2004[McLaren, V. L., Kirk, C. A., Poisot, M., Castellanos, M. & West, A. R. (2004). Dalton Trans. 19, 3042-3047.]). The positions of the Li atoms are (1/4, 1/4, 1/4), (3/4, 1/4, 1/4), (3/4, 3/4, 1/4), (1/4, 3/4, 1/4), (1/4, 1/4, 1/2), (3/4, 1/4, 1/2), (3/4, 3/4, 1/2), (1/4, 3/4, 1/2), as shown in Fig. 3[link](b). Therefore, the structure factor of Li0.5Ni0.5O along [1[\overline{1}]1] changes to

[|F(1\overline{1}1)^{\rm cal}_{\rm Li_{0.5}Ni_{0.5}O}|^2 = 16[0.5f(1\overline{1}1)_{\rm Ni}-f(1\overline{1}1)_{\rm O}]^2.\eqno(11)]

Let us introduce B3 for all the Li atoms sitting in interstitial sites. The value is 3.1 (15) Å 2 using equations (8)[link] and (11)[link], as shown in Table 2[link]. The difference between B3 for the Li interstitial model and B1 is smaller than that between B2 for the Li substitute model and B1. This implies that some Li atoms are probably located in the interstitial sites in a heavily doped Li0.5Ni0.5O thin film. These results help us to understand the physical properties of LixNi1-xO to develop LixNi1-xO-based devices.

Table 2
B2 and B3 factors, Debye temperature θM, r.m.s. amplitude of atomic vibration ([\overline{u^2}^{1/2}]), ([1 \overline 1 1]) lattice spacing [d_{1 \overline 1 1}], and ratio [r = \overline{u^2}^{1/2} / d_{1 \overline 1 1}] of Li0.5Ni0.5O thin films at 298 K

Site B2 or B32) θM (K) [\overline{u^2}^{1/2}] (Å) [d_{1\overline{1}1}] (Å) r (%)
Substitutional 6.5 (15) 147 (22) 0.29 (3) 2.368 (4) 12.2 (13)
Interstitial 3.1 (15) 213 (45) 0.20 (6) 2.368 (4) 8.4 (25)

6. Concluding remarks

In conclusion, the conventional B1 factor of an Li0.5Ni0.5O thin film was estimated to be 1.8 (4) Å2 using X-ray diffraction by considering the 1[\overline{1}]1 and 2[\overline{2}]2 reflections. The corresponding Debye temperature was 281 (39) K. Furthermore, the pseudo-Debye–Waller factor of the Li0.5Ni0.5O thin film was obtained using the intensity ratio between the 1[\overline{1}]1 Bragg peaks of Li0.5Ni0.5O and NiO thin films. The atomic disorder parameter that we proposed was evaluated to be 0.66. The disorder may include combined effects of thermal vibration, interstitial atoms and defects caused by Li doping. Interstitial Li atoms could be present in heavily Li-doped Li0.5Ni0.5O thin films because the proposed B3 factor was smaller than B2 determined for a film with Li ions only in Ni sites.

Acknowledgements

The thin-film X-ray diffraction measurements were approved by the NIMS Synchrotron X-ray Station at SPring-8 under proposal Nos. 2011B4508, 2012A4505, 2012B4505 and 2014B4900. This work was also partly supported by the Ministry of Education, Culture, Sports, Science and Technology of Japan [OS: Grants-in-Aid for Basic Research (C) No. 15K04616]. The authors are grateful to Dr K. Yokoyama, Dr S. Takeda, Professor Y. Kagoshima and Professor J. Matsui, University of Hyogo, for their technical contribution.

References

First citationBorca, C. N., Komesu, T., Jeong, H. K., Dowben, P. A., Ristoiu, D., Hordequin, C., Pierre, J. & Nozières, J. P. (2000). Appl. Phys. Lett. 77, 88–90.  CrossRef CAS Google Scholar
First citationBrown, P. J., Fox, A. G., Maslen, E. N., O'Keefe, M. A. & Willis, B. T. M. (2006). International Tables for Crystallography, Vol. C, Mathematical, Physical and Chemical Tables, 1st online ed., edited by E. Prince, ch. 6.1, pp. 554–595. Chester: International Union of Crystallography.  Google Scholar
First citationChawla, S., Jayanthi, K. & Kotnala, R. K. (2009). Phys. Rev. B, 79, 125204.  CrossRef Google Scholar
First citationChipman, D. R. & Batterman, B. W. (1963). J. Appl. Phys. 34, 912–913.  CrossRef CAS Google Scholar
First citationDutta, T., Gupta, P., Gupta, A. & Narayan, J. (2010). J. Appl. Phys. 108, 083715.  CrossRef Google Scholar
First citationFreer, R. (1981). J. Mater. Sci. 16, 3225–3227.  CrossRef CAS Web of Science Google Scholar
First citationGarduno-Wilches, I. & Alonso, J. C. (2013). Int. J. Hydrogen Energy, 38, 4213–4219.  CAS Google Scholar
First citationGazzara, C. P. & Middleton, R. M. (1964). Adv. X-ray Anal. 7, 14–16.  CAS Google Scholar
First citationGoodenough, J. B., Wickham, D. G. & Croft, W. J. (1958). J. Phys. Chem. Solids, 5, 107–116.  CrossRef CAS Google Scholar
First citationGuo, W., Hui, K. N. & Hui, K. S. (2013). Mater. Lett. 92, 291–295.  CrossRef CAS Google Scholar
First citationHerbstein, F. H. (1961). Adv. Phys. 10, 313–355.  CrossRef CAS Web of Science Google Scholar
First citationJames, R. W. (1962). The Optical Principles of the Diffraction of X-rays. Woodbridge: Ox Bow Press.  Google Scholar
First citationKulkarni, R. G. & Bichile, G. K. (1977). J. Appl. Phys. 48, 1844–1847.  CrossRef CAS Google Scholar
First citationLonsdale, K. (1948). Acta Cryst. 1, 142–149.  CrossRef IUCr Journals Google Scholar
First citationMatsumuro, A., Kobayashi, M., Kikegawa, T. & Senoo, M. (1990). J. Appl. Phys. 68, 2719–2722.  CrossRef CAS Google Scholar
First citationMcLaren, V. L., Kirk, C. A., Poisot, M., Castellanos, M. & West, A. R. (2004). Dalton Trans. 19, 3042–3047.  CrossRef PubMed Google Scholar
First citationMoulki, H., Park, D. H., Min, B.-K., Kwon, H., Hwang, S.-J., Choy, J.-H., Toupance, T., Campet, G. & Rougier, A. (2012). Electrochim. Acta, 74, 46–52.  CrossRef CAS Google Scholar
First citationMurakami, T. (1953). J. Phys. Soc. Japan, 8, 36–40.  CrossRef CAS Google Scholar
First citationOhta, H., Hirano, M., Nakahara, K., Maruta, H., Tanabe, T., Kamiya, M., Kamiya, T. & Hosono, H. (2003). Appl. Phys. Lett. 83, 1029–1031.  CrossRef CAS Google Scholar
First citationPalatnikov, M. N., Biryukova, I. V., Sidorov, N. V., Denisov, A. V., Kalinnikov, V. T., Smith, P. G. R. & Shur, V. Ya. (2006). J. Cryst. Growth, 291, 390–397.  CrossRef CAS Google Scholar
First citationPathak, P. D. & Trivedi, J. M. (1973). Acta Cryst. A29, 45–49.  CrossRef IUCr Journals Google Scholar
First citationSakata, O. & Hashizume, H. (1997). Acta Cryst. A53, 781–788.  CrossRef CAS IUCr Journals Google Scholar
First citationSakata, O., Yi, M. S. Matsuda, A., Liu, J., Sato, S., Akiba, S., Sasaki, A. & Yoshimoto, M. (2004). Appl. Surf. Sci. 221, 450–454.  CrossRef CAS Google Scholar
First citationSasaki, T., Oguchi, T. & Katayama-Yoshida, H. (1993). Physica B, 185, 159–163.  CrossRef CAS Google Scholar
First citationShiraishi, N., Kato, Y., Arai, H., Tsuchimine, N., Kobayashi, S., Mitsuhashi, M., Soga, M., Kaneko, S. & Yoshimoto, M. (2010). Jpn. J. Appl. Phys. 49, 108001.  CrossRef Google Scholar
First citationWathore, T. N. & Kulkarni, R. G. (1980). J. Mater. Sci. 15, 2927–2930.  CrossRef CAS Google Scholar
First citationYamauchi, R., Tan, G., Shiojiri, D., Koyama, K., Kaneko, S., Matsuda, A. & Yoshimoto, M. (2012). Jpn. J. Appl. Phys. 51, 06FF16.  CrossRef Google Scholar
First citationYang, A. L., Sakata, O., Yamauchi, R., Katsuya, Y., Kumara, L. S. R., Shimada, Y., Matsuda, A. & Yoshimoto, M. (2014). Appl. Surf. Sci. 320, 787–790.  CrossRef CAS Google Scholar
First citationYoshimoto, M., Maeda, T., Ohnishi, T., Koinuma, H., Ishiyama, O., Shinohara, M., Kubo, M., Miura, R. & Miyamoto, A. (1995). Appl. Phys. Lett. 67, 2615–2617.  CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Volume 48| Part 6| December 2015| Pages 1896-1900
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds