We report the fabrication of atomically abrupt interfaces of titanium dihydride (δ-TiH2) films and α-Al2O3(001) substrates. With the assistance from reactive hydrogen in plasma, single-phase δ-TiH2 epitaxial thin films were grown on α-Al2O3(001) substrates using the reactive magnetron sputtering technique. Scanning transmission electron microscopy measurements revealed an atomically abrupt interface at the δ-TiH2(111) film and Al2O3(001) substrate. These results indicate that the reactive magnetron sputtering has great potential to deposit various epitaxial thin films of hydrides restricted by the hydrogenation limit. The fabrication of high-quality hydride epitaxial thin films with atomically controlled interfaces paves the way for future hydride electronics.

Metal hydrides have been extensively studied as hydrogen storage materials for practical applications such as hydrogen fuel systems and fuel cells.1,2 In addition, the recent reports in hydrides on ionic conduction of metal (Li+, Na+, Mg2+) cations3 and hydride (H) ion4,5 have paved the way for electrochemical applications. Furthermore, high transition-temperature superconductivity in H3S6 and PdHx7 has attracted enormous attention in the field of condensed-matter physics. Accordingly, metal hydrides are promising materials in a wide variety of fields ranging from energy devices to electronics.8 

For the applications of metal hydrides in electronics,9 atomically abrupt interfaces between the substrate and film and/or between the films in multilayer-structured devices are essential. Since exotic physical properties such as high-mobility two-dimensional electron gases, magnetism, and superconductivity appear at the heterointerfaces of oxides,10 “heterointerfaces of hydrides” have the great potential to emerge properties using strain, superlattice, and field-effect. For the fabrication of metal-hydride epitaxial thin films and abrupt interfaces in a well-controlled fashion, the direct growth of single-phase epitaxial thin films is required. Until now, there have been only a few reports on the fabrication of metal-hydride epitaxial thin films.9,11–13 However, epitaxial thin films of LiH11 and TiH212 directly grown using pulsed laser deposition contained impurity phases of Li2O and metallic Ti, respectively. Accordingly, it is of significant importance to introduce another deposition technique for obtaining single-phase metal hydride epitaxial thin films. For this purpose, reactive magnetron sputtering is a promising candidate. The growth of polycrystalline ZrH2 thin films using reactive magnetron sputtering has been reported,14 implying the potential of the sputtering technique to obtain single-phase metal hydride epitaxial thin films and atomically controlled interfaces.

Here, we demonstrate the fabrication of the atomically abrupt interface of single-phase epitaxial titanium dihydride (δ-TiH2, fluorite-type, Fm−3m, a = 0.446 nm15) thin films and α-Al2O3 (001) substrates. We adopted the magnetron sputtering technique because we expected the assistance of hydrogenation for the thin films from reactive hydrogen in plasma, enabling the direct growth of metal hydride thin films without impurity phases. The (111)-oriented δ-TiH2 epitaxial films were successfully fabricated at a substrate temperature (Ts) of 150°C and a hydrogen partial pressure (pH2) of 0.05 Pa. Scanning transmission electron microscopy measurements revealed that the δ-TiH2 (111)/Al2O3 (001) interface is atomically abrupt. Furthermore, the temperature dependence of resistivity shows metallic behavior with no upturn over the temperature range of 4–300 K, showing that impurity phases are not present in the films. Our study shows that reactive magnetron sputtering has an advantage in fabricating atomically controlled interfaces of single-phase hydride epitaxial thin films and substrates.

Thin films of δ-TiH2 were deposited on α-Al2O3 (001) (Shinkosha Corp.) substrates using reactive magnetron sputtering. A Ti metal plate (purity: 99.99%, 2 in. diameter, Toshima Manufacturing Co., Ltd.) was used as a sputtering target, and a mixture of pure Ar and H2 gases was introduced into a vacuum chamber (background pressure: ∼5 × 10−5 Pa) for the sputtering. During the depositions, the Ar partial pressure was fixed at 1.0 Pa with a constant flow rate of 10 SCCM in the growth chamber, and pH2 was varied with different flow rates. The dc power supply at the Ti target was maintained at 50 W. The Ts values were varied from room temperature (RT) to 300°C. The typical film thickness was 90 nm with a growth time of 30 min. The thin-film deposition rate was 180 nm/h. As the H2 partial pressure in the sputtering process was increased, the deposition rate decreased: 200 nm/h and 180 nm/h in Ar and Ar/H2 (5%) atmosphere, respectively, suggesting that the Ti target surface was hydrogenated during the deposition. The structural properties of the thin films were characterized using x-ray diffraction (XRD, Bruker D8 DISCOVER). Annular dark field-scanning transmission electron microscopy (ADF-STEM) observations and selected area diffraction pattern (SADP) analyses were conducted with JEM-ARM200F (JEOL Co., Ltd.) operated at 200 kV. The thin foil for STEM observations was prepared with the SMF2000 (Hitachi High-Tech Science Co.) focused ion beam system. Electron transport measurements were carried out in a typical Hall-bar geometry using a Gifford-McMahon refrigerator with superconducting magnets (Niki Glass Co., Ltd., LTS205D-CM9T-TL-VT).

Figure 1(a) shows the obtained XRD results of the Ts dependence (pH2 = 0.15 Pa) of the thin films deposited on Al2O3 (001) substrates. The out-of-plane XRD scan of the thin film deposited at Ts = 300°C shows a 002 peak of the metallic α-Ti (hexagonal, 2θ = 38.422°16). As Ts was decreased to 200°C, the α-Ti 002 peak diminished, and a peak appeared at 2θ ∼ 35.0°, corresponding to δ-TiH2 111.15 We here stress that the α-Ti 002 peak completely disappeared in the films deposited at Ts 200°C (even at RT). Thus, Ti metal was fully hydrogenated in films deposited in a reactive magnetron sputtering process at low Ts, indicating the positive effect on the hydrogenation of films.

FIG. 1.

X-ray diffraction (XRD) structural characterization of films deposited using reactive magnetron sputtering. (a) Substrate temperature (Ts) dependence of the out-of-plane XRD scans. (b) Hydrogen partial pressure (pH2) dependence of the out-of-plane XRD scans. [(c) and (d)] Pole figure measurements for (c) δ-TiH2 200 and (d) Al2O3 113. (e) Schematic illustration of the epitaxial relationship between the (111)-oriented δ-TiH2 film and Al2O3 (001) substrate.

FIG. 1.

X-ray diffraction (XRD) structural characterization of films deposited using reactive magnetron sputtering. (a) Substrate temperature (Ts) dependence of the out-of-plane XRD scans. (b) Hydrogen partial pressure (pH2) dependence of the out-of-plane XRD scans. [(c) and (d)] Pole figure measurements for (c) δ-TiH2 200 and (d) Al2O3 113. (e) Schematic illustration of the epitaxial relationship between the (111)-oriented δ-TiH2 film and Al2O3 (001) substrate.

Close modal

We next examined the pH2 dependence (Ts = 150°C) of the δ-TiH2 thin films. Figure 1(b) shows the out-of-plane XRD patterns of the films deposited at pH2 = 0 (no additional H2 gas), 0.01, 0.02, 0.03, 0.05, and 0.15 Pa. Under pH2 = 0 Pa, only the α-Ti 002 peak is clearly visible. As the pH2 is increased, the intensity of the α-Ti 002 peak became smaller, and a δ-TiH2 111 peak emerged. Further increase in the pH2 value to 0.05 Pa led to the deposition of single-phase δ-TiH2 (111) thin films. Note that the intensity of the δ-TiH2 111 peak in pH2 = 0.15 Pa was smaller than those in pH2 = 0.03 and 0.05 Pa. This is because polycrystalline films were obtained in extremely high pH2 (not shown).

To study the in-plane crystal orientation of the δ-TiH2 films, pole figure measurements were performed for δ-TiH2 200 (2θ = 40.41°, χ = 54.7°) and Al2O3 113 (2θ = 43.36°, χ = 61.2°). The pole figure measurements indicate that the peaks of TiH2 200 and Al2O3 113 appear at the same φ-angles, [Figs. 1(c) and 1(d)], demonstrating the epitaxial growth of TiH2 thin films. Note that the six diffraction spots for δ-TiH2 200 indicate existence of two rotational domains. The obtained δ-TiH2 films were relaxed, possibly due to the large in-plane lattice mismatch of 14.2%: δ-TiH2 (0.314 nm for Ti–Ti) to Al2O3 (0.275 nm for O–O). The in-plane epitaxial relationship of [112¯]TiH2 // [110]Al2O3, which is consistent with a previous report using pulsed laser deposition.12 

The extent of hydrogenation and the crystallinity in the deposited films as a function of Ts and pH2 are investigated. To evaluate the extent of hydrogenation, we here introduce the δ-TiH2 fraction as

STiH2_111/  STiH2_111+STi_002,

where STiH2_111 and STi_002 are the areas of the XRD peak at δ-TiH2 111 and α-Ti 002, respectively. Figure 2(a) is a color gradation plot of the δ-TiH2 fraction, as a function of Ts (horizontal) and pH2 (vertical). Single-phase δ-TiH2 films (δ-TiH2 fraction = 1) are shown in the red regions, while a metallic α-Ti appears in the blue regions, and accordingly, lower Ts and higher pH2 lead to a δ-TiH2 phase. In Fig. 2(b), we also show a color map of a rocking curve full-width at half-maximum (FWHM) obtained at δ-TiH2 111 (2θ = 34.8°). Red regions indicate the δ-TiH2 (111) epitaxial films with high crystallinity. Interestingly, higher pH2 led to the formation of polycrystalline films, indicating that extremely high pH2 hinders the fabrication of high-quality δ-TiH2 epitaxial films [Fig. 1(b)]. According to these two plots, the best conditions for the δ-TiH2 (111) epitaxial growth are determined as Ts = 150°C and pH2 = 0.05 Pa. The smallest rocking curve FWHM value obtained is 0.67°, which is smaller than the previously reported value (1.4°).12 

FIG. 2.

Growth phase diagram of the fabricated δ-TiH2 films as a function of the substrate temperature (Ts, horizontal) and hydrogen partial pressure (pH2, vertical). Black dots show the experimental deposition conditions for 30 films. (a) Color gradation map of the extent of hydrogenation (see main text for the definitions) characterized by the XRD peak area. Red and blue regions correspond to δ-TiH2 and α-Ti, respectively. (b) Color gradation map of the full-width at half maximum (FWHM) of the TiH2 111 peak. Red regions indicate higher crystallinity.

FIG. 2.

Growth phase diagram of the fabricated δ-TiH2 films as a function of the substrate temperature (Ts, horizontal) and hydrogen partial pressure (pH2, vertical). Black dots show the experimental deposition conditions for 30 films. (a) Color gradation map of the extent of hydrogenation (see main text for the definitions) characterized by the XRD peak area. Red and blue regions correspond to δ-TiH2 and α-Ti, respectively. (b) Color gradation map of the full-width at half maximum (FWHM) of the TiH2 111 peak. Red regions indicate higher crystallinity.

Close modal

We here demonstrate an atomically abrupt interface of δ-TiH2 (111) epitaxial thin films and Al2O3 (001) substrates using cross-sectional ADF-STEM imaging. In a wide-view ADF-STEM image [Fig. 3(a)], the δ-TiH2 epitaxial film shows the uniform contrast. We could observe an additional TiOx layer on the top of δ-TiH2, possibly formed during the deposition of an Os protective layer. The layer was deposited using plasma-assisted chemical vapor deposition with OsOx precursor, prior to the specimen preparation for STEM measurements. The electron diffraction patterns obtained in the film region show sharp spots reflecting (111)-oriented fluorite structures with two rotational domains [Figs. 3(b) and 3(c)], consistent with the XRD results. In a close-up view near the TiH2/Al2O3 interface [Fig. 3(d)], we clearly visualize the ordering of the Ti atoms in the film regions. In the out-of-plane direction, the ordering is (111)-oriented fcc stacking (ABCABC…) corresponding to Ti sites formed in the fluorite structures of TiH2, distinct from the metallic phases of α-Ti (hcp) and β-Ti (bcc). Consequently, an atomically abrupt interface can be achieved even in a hydride film, illustrating the great potential of magnetron sputtering for the development of hydride electronics in future.

FIG. 3.

Atomic-scale structural imaging of a TiH2/Al2O3 (001) epitaxial film fabricated at Ts = 150°C and pH2 = 0.05 Pa using annular dark field-scanning transmission electron microscopy (ADF-STEM). (a) A wide-view ADF-STEM image. [(b) and (c)] Electron diffraction patterns obtained on the (b) TiH2 film and (c) Al2O3 substrate. Note that two domains are observed in the diffraction patterns obtained in (b). (d) A close-up ADF-STEM image near the interface at the TiH2 (111) epitaxial film and Al2O3 (001) substrate. Clear fcc-like (ABCABC…) stacking is observed in the film region.

FIG. 3.

Atomic-scale structural imaging of a TiH2/Al2O3 (001) epitaxial film fabricated at Ts = 150°C and pH2 = 0.05 Pa using annular dark field-scanning transmission electron microscopy (ADF-STEM). (a) A wide-view ADF-STEM image. [(b) and (c)] Electron diffraction patterns obtained on the (b) TiH2 film and (c) Al2O3 substrate. Note that two domains are observed in the diffraction patterns obtained in (b). (d) A close-up ADF-STEM image near the interface at the TiH2 (111) epitaxial film and Al2O3 (001) substrate. Clear fcc-like (ABCABC…) stacking is observed in the film region.

Close modal

Finally, we performed electron transport measurements using a Hall-bar geometry on δ-TiH2 (111) epitaxial films grown at Ts = 150°C and pH2 = 0.05 Pa. The temperature dependence of the longitudinal resistivity [ρxx(T)] shows metallic behavior over the entire temperature range of 4–300 K [Fig. 4(a)]. The residual resistance ratio [ρxx(300 K)/ρxx(4 K)] is 2.4. Note that the upturn at around 15 K reported in the previous study12 did not appear, suggesting that an impurity phase is not present in our films. Figure 4(b) shows the magnetic field dependence of the transverse resistance (Rxy) obtained at 4 K. The slope in a linear response gives a Hall coefficient of −9.5 × 10−5 (cm3/C), and thus, the carrier concentration and mobility at 4 K are obtained as 6.6 × 1022 cm−3 and 2.3 cm2 V−1 s−1, respectively. Given that the valence of Ti is 2+ (d2 system), the obtained carrier concentration of 6.6 × 1022 cm−3 is consistent with the order of the Ti atom density (4.5 × 1022 cm−3).

FIG. 4.

Electron transport properties of a TiH2/Al2O3 (001) film grown at Ts = 150°C and pH2 = 0.05 Pa. (a) Temperature dependence of the longitudinal resistivity (ρxx). The residual resistance ratio [RRR, ρxx(300 K)/ρxx(4 K)] of ∼2.4 is obtained. (b) Magnetic-field dependence of transverse resistance (Rxy). A clear linear dependence was observed within the range of ±5 T.

FIG. 4.

Electron transport properties of a TiH2/Al2O3 (001) film grown at Ts = 150°C and pH2 = 0.05 Pa. (a) Temperature dependence of the longitudinal resistivity (ρxx). The residual resistance ratio [RRR, ρxx(300 K)/ρxx(4 K)] of ∼2.4 is obtained. (b) Magnetic-field dependence of transverse resistance (Rxy). A clear linear dependence was observed within the range of ±5 T.

Close modal

In conclusion, we investigated the epitaxial growth of single-phase δ-TiH2 (111) thin films on Al2O3 (001) substrates using reactive magnetron sputtering. We studied the growth conditions of δ-TiH2 epitaxial films as a function of Ts and pH2. As a result, we successfully fabricated δ-TiH2 epitaxial films at Ts = 150°C and pH2 = 0.05 Pa, whereas high pH2 and low Ts resulted in polycrystalline films. Atomic-resolution STEM measurements clearly visualized fcc-like Ti stacking based on fluorite structures of TiH2 and an atomically abrupt interface of TiH2 and Al2O3. Furthermore, transport measurements indicate metallic behavior even at low temperatures, consistent with the result that no impurity phase is observed in the XRD and STEM. Our present study indicates that reactive magnetron sputtering is advantageous to fabricate high-quality metal hydride epitaxial films and abrupt interfaces for future hydride electronics.

This research was supported by the World Premier International Research Center Initiative (WPI Initiative) of the Ministry of Education, Culture, Sports, Science and Technology of Japan, JST-CREST (JPMJCR1523) and JST-ALCA. R.S. acknowledges Kakenhi (No. 17H05216). T.H. acknowledges Kakenhi (Nos. 26246022 and 16K14088). H.O. acknowledges Kakenhi (Nos. 15H05546 and 15K14155) and the TEPCO Memorial Foundation. S.O. acknowledges Kakenhi (No. 25220911). This work was performed under the Inter-university Cooperative Research Program of the Institute for Materials Research, Tohoku University (Proposal Nos. 16K0025 and 17K0060). STEM observations were carried out at the National Institute for Materials Science (NIMS) Battery Research Platform. We thank Professor Tsutomu Nojima for his help in designing our custom-made GM fridge instruments. We also thank Professor Akira Ohtomo, Dr. Kohei Yoshimatsu, and Mr. Takuto Soma for their support of electron transport measurements.

1.
L.
Schlapbach
and
A.
Züttel
, “
Hydrogen-storage materials for mobile applications
,”
Nature
414
,
353
(
2001
).
2.
S.
Orimo
,
Y.
Nakamori
,
J. R.
Eliseo
,
A.
Züttel
, and
C. M.
Jensen
, “
Complex hydrides for hydrogen storage
,”
Chem. Rev.
107
,
4111
(
2007
).
3.
A.
Unemoto
,
M.
Matsuo
, and
S.
Orimo
, “
Complex hydrides for electrochemical energy storage
,”
Adv. Funct. Mater.
24
,
2267
(
2014
).
4.
M. C.
Verbraeken
,
C.
Cheung
,
E.
Suard
, and
J. T. S.
Irvine
, “
High H ionic conductivity in barium hydride
,”
Nat. Mater.
14
,
95
(
2015
).
5.
G.
Kobayashi
,
Y.
Hinuma
,
S.
Matsuoka
,
A.
Watanabe
,
M.
Iqbal
,
M.
Hirayama
,
M.
Yonemura
,
T.
Kamiyama
,
I.
Tanaka
, and
R.
Kanno
, “
Pure H conduction in oxyhydrides
,”
Science
351
,
1314
(
2016
).
6.
A. P.
Drozdov
,
M. I.
Eremets
,
I. A.
Troyan
,
V.
Ksenofontov
, and
S. I.
Shylin
, “
Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system
,”
Nature
525
,
73
(
2015
).
7.
H. M.
Syed
,
T. J.
Gould
,
C. J.
Webb
,
E.
Mac
, and
A.
Gray
, “
Superconductivity in palladium hydride and deuteride at 52-61 Kelvin
,” e-print arXiv:1608.01774.
8.
R.
Mohtadi
and
S.
Orimo
, “
The renaissance of hydrides as energy materials
,”
Nat. Rev. Mater
2
,
16091
(
2016
).
9.
H.
Oguchi
,
T.
Ikeshoji
,
T.
Ohsawa
,
S.
Shiraki
,
H.
Kuwano
,
S.
Orimo
, and
T.
Hitosugi
, “
Epitaxial thin film growth of LiH using a liquid-Li atomic template
,”
Appl. Phys. Lett.
105
,
211601
(
2014
).
10.
H. Y.
Hwang
,
Y.
Iwasa
,
M.
Kawasaki
,
B.
Keimer
,
N.
Nagaosa
, and
Y.
Tokura
, “
Emergent phenomena at oxide interfaces
,”
Nat. Mater.
11
,
103
(
2012
).
11.
H.
Oguchi
,
S.
Isobe
,
H.
Kuwano
,
S.
Shiraki
,
S.
Orimo
, and
T.
Hitosugi
, “
Pulsed laser deposition of air-sensitive hydride epitaxial thin films: LiH
,”
APL Mater.
3
,
096106
(
2015
).
12.
K.
Yoshimatsu
,
T.
Suzuki
,
N.
Tsuchimine
,
K.
Horiba
,
H.
Kumigashira
,
T.
Oshima
, and
A.
Ohtomo
, “
Direct growth of metallic TiH2 thin films by pulsed laser deposition
,”
Appl. Phys. Express
8
,
035801
(
2015
).
13.
J.
Hayoz
,
Th.
Pillo
,
M.
Bovet
,
A.
Züttel
,
St.
Guthrie
,
G.
Pastore
,
L.
Schlapbach
, and
P.
Aebi
, “
Preparation and characterization of clean, single-crystalline YHx films (0 ≤ x ≤ 2.9) on W(110)
,”
J. Vac. Sci. Technol., A
18
,
2417
(
2000
).
14.
H.
Högberg
,
L.
Tengdelius
,
M.
Samuelsson
,
F.
Eriksson
,
E.
Broitman
,
J.
Lu
,
J.
Jensen
, and
L.
Hultman
, “
Reactive sputtering of δ-ZrH2 thin films by high power impulse magnetron sputtering and direct current magnetron sputtering
,”
J. Vac. Sci. Technol., A
32
,
041510
(
2014
).
15.
S. K.
Dolukhanyan
, “
Synthesis of novel compounds by hydrogen combustion
,”
J. Alloy Compd.
253-254
,
10
(
1997
).
16.
Powder diffraction file: No. 44–1294.