ArticlePDF Available

Electronic Properties and Bonding in ZrHx Thin Films Investigated by Valence-Band X-ray Photoelectron Spectroscopy

Authors:
  • IHI Ionbond

Abstract and Figures

The electronic structure and chemical bonding in reactively magnetron sputtered ZrH x (x = 0.15, 0.30, 1.16) thin films with oxygen content as low as 0.2 at.% are investigated by 4d valence band, shallow 4p core-level, and 3d core-level x-ray photoelectron spectroscopy. With increasing hydrogen content, we observe significant reduction of the 4d valence states close to the Fermi level as a result of redistribution of intensity toward the H 1s-Zr 4d hybridization region at ∼6 eV below the Fermi level. For low hydrogen content (x = 0.15, 0.30), the films consist of a superposition of hexagonal closest-packed metal (α phase) and understoichiometric δ-ZrH x (CaF 2-type structure) phases, while for x = 1.16, the films form single-phase ZrH x that largely resembles that of stoichiometric δ-ZrH 2 phase. We show that the cubic δ-ZrH x phase is metastable as thin film up to x = 1.16, while for higher H contents the structure is predicted to be tetragonally distorted. For the investigated ZrH 1.16 film, we find chemical shifts of 0.68 and 0.51 eV toward higher binding energies for the Zr 4p 3/2 and 3d 5/2 peak positions, respectively. Compared to the Zr metal binding energies of 27.26 and 178.87 eV, this signifies a charge transfer from Zr to H atoms. The change in the electronic structure, spectral line shapes, and chemical shifts as a function of hydrogen content is discussed in relation to the charge transfer from Zr to H that affects the conductivity by charge redistribution in the valence band.
Content may be subject to copyright.
PHYSICAL REVIEW B 96, 195103 (2017)
Electronic properties and bonding in ZrHxthin films investigated by valence-band
x-ray photoelectron spectroscopy
Martin Magnuson,*Susann Schmidt, Lars Hultman, and Hans Högberg
Thin Film Physics Division, Department of Physics, Chemistry and Biology (IFM), Linköping University, SE-58183 Linköping, Sweden
(Received 24 May 2017; revised manuscript received 20 September 2017; published 1 November 2017)
The electronic structure and chemical bonding in reactively magnetron sputtered ZrHx(x=0.15, 0.30, 1.16)
thin films with oxygen content as low as 0.2 at.% are investigated by 4d valence band, shallow 4p core-level,
and 3d core-level x-ray photoelectron spectroscopy. With increasing hydrogen content, we observe significant
reduction of the 4d valence states close to the Fermi level as a result of redistribution of intensity toward the H
1s–Zr 4d hybridization region at 6 eV below the Fermi level. For low hydrogen content (x=0.15, 0.30), the
films consist of a superposition of hexagonal closest-packed metal (αphase) and understoichiometric δ-ZrHx
(CaF2-type structure) phases, while for x=1.16, the films form single-phase ZrHxthat largely resembles that
of stoichiometric δ-ZrH2phase. We show that the cubic δ-ZrHxphase is metastable as thin film up to x=1.16,
while for higher H contents the structure is predicted to be tetragonally distorted. For the investigated ZrH1.16
film, we find chemical shifts of 0.68 and 0.51 eV toward higher binding energies for the Zr 4p3/2and 3d5/2
peak positions, respectively. Compared to the Zr metal binding energies of 27.26 and 178.87 eV, this signifies
a charge transfer from Zr to H atoms. The change in the electronic structure, spectral line shapes, and chemical
shifts as a function of hydrogen content is discussed in relation to the charge transfer from Zr to H that affects
the conductivity by charge redistribution in the valence band.
DOI: 10.1103/PhysRevB.96.195103
I. INTRODUCTION
Zirconium hydrides are important in many applications
in the nuclear industry [1], in getter vacuum pumps, as
hydrogen storage, in powder metallurgy, and as zircaloy [2].
The hardness, ductility, and tensile strength of ZrHxalloys can
be controlled by varying the hydrogen content [3]. Increasing
the H content results in substoichiometric δ-ZrHxphase
(CaF2-type structure, x=∼1.62[4] as shown in Fig. 1)
and a body-centered tetragonal εphase (x=1.752) with
aThH
2-type structure [5] has also been observed [4]. The
δ-ZrHxstructure originates from incerting H atoms to occupy
all or parts of the tetrahedral interstitials in the CaF2structure.
Hydrogen acts as a hardening element that prevents dislocation
movements and the hydride material becomes a ceramic that
is harder, but less ductile than Zr metal [6].
Traditionally, bulk Zr hydrides are synthesized by
annealing Zr metal in hydrogen gas for a time period of days
to a few weeks at high temperatures between 400 and 900 C
for a homogeneous diffusion process at different H contents
[7]. However, defects in the bulk of the Zr metal result in
variations in the diffusion rate for hydrogen that cause ZrHx
alloys with composition gradients and less well-defined
polycrystalline structure including grain boundaries. At room
temperature in air, ZrHxquickly forms a nanometer-thin
surface oxide layer that prevents further oxidation into the
bulk. With increased annealing temperature, oxygen proceeds
deeper into the bulk of the material, in particular along grain
boundaries [8], and a few percent of oxygen cannot be avoided
with the annealing-diffusion hydration synthesis method.
Consequently, the above conditions hinder adequate bonding
determination in hydrides by spectroscopy.
Previous experiments on bulk ZrHxusing ultraviolet pho-
toelectron spectroscopy showed significant changes in the
*Corresponding author: Martin.Magnuson@liu.se
electronic structure for high hydrogen contents (x=
1.631.94) both at the states near the Fermi level (EF) and
in deeper-lying states around 7eV[9]. These changes were
associated with a phase transition from a cubic (fcc) structure
to a face-centered tetragonal (fct) structure as a result of
spontaneous symmetry breaking that removes the degeneracy
by a distortion known as the Jahn-Teller effect [10] that lowers
the total energy of the 4d25s2electron configuration in the Zr
valence band.
For high hydrogen contents (x=1.631.9), core-level
x-ray photoelectron spectroscopy (XPS) studies of bulk ZrHx
materials indicated a significant shift of the Zr 4p and 3d core
levels by 1.0 and 0.7 eV [7], respectively. For x=1.521.68,
valence-band XPS also showed an interstitial Zr–H bonding
peak around 6.4 eV below EF[11]. Although these studies
indicate significant changes in the Zr–H bond for different
hydrogen contents, the spectra are affected by superimposed
O2pstates that occur in the same energy region (5–8 eV) as
the Zr–H bonding structures [12].
For the ZrHxthin films in this work, we apply XPS at the
Zr 3d,4p core levels and the 4d valence band to investigate
the electronic structures to determine the chemical bonding
and conductivity properties as a function of relatively low
hydrogen content (x=0.15, 0.30, and 1.16) and compare the
result to bulk α-Zr as reference. These hydride compositions
were chosen to obtain a spread in the type and amount
of Zr–H bonds. Deposited δ-ZrHxfilms with CaF2-type
structure are found to be stable outside the homogeneity
region determined for bulk δ-ZrH2with compositions ranging
from about ZrH1.5to stoichiometric [13]. To control the
structure and hence the electronic properties of the materials
and to overcome the problems with oxidation, we deposited
homogeneous thin films of ZrHxusing reactive direct current
magnetron sputtering (rDCMS). The thin films were grown
without external heating and contamination of oxygen, and
other contaminates were very low. Thin films are particularly
2469-9950/2017/96(19)/195103(7) 195103-1 ©2017 American Physical Society
MAGNUSON, SCHMIDT, HULTMAN, AND HÖGBERG PHYSICAL REVIEW B 96, 195103 (2017)
FIG. 1. Unit cell of stoichiometric δ-ZrH2(space group 225) with
Zr-atoms (red spheres) in (0,0,0), (½,½,0), (½,0,½), and (0,½,½)and
with H (white spheres) in all tetrahedral sites (¾,¼,¼), (¼,¾,¼),
(¼,¼,¾), (¾,¾¾), (¼,¼,¼), (¾,¾,¼), (¼,¾,¾), and (¾,¼,¼).
favorable for correct characterization of the material’s
fundamental bonding properties and to identify different
phases in understoichiometric transition-metal hydrides.
II. EXPERIMENTAL
A. Thin-film deposition and XRD characterization
The Zr-H films studied were deposited on Si(100) sub-
strates by rDCMS. ZrHxfilms with x=0.15, 0.30, and 1.16
were deposited in an industrial high-vacuum coating system
(CemeCon AG, Würselen, Germany). Here, a zirconium target
wassputteredinAr(99.9997%)/H2(99.9996% mixtures using
a fixed Ar partial pressure of 0.42 Pa with 5, 10, and 20% H2.
The substrate bias was set to 80 V and no external substrate
heating was applied. The films were deposited for 120 s,
resulting in films with thicknesses ranging between 800
and 840 nm. More detailed information on the deposition
conditions is presented in Ref. [14]. The investigated α-Zr
reference sample was a commercial zirconium target with a
purity of 99.9% (Kurt J. Lesker Company, Clairton, PA, USA).
B. XPS measurements
Valence-band (VB) and XPS measurements of the Zr 3d,
4p,4dvalence band and O 1score-level regions were per-
formed in a surface analysis system (AXIS UltraDLD, Kratos,
Manchester, UK) using monochromatic Al-Kα(1486.6 eV)
radiation with an incidence angle of 54and a spot size
of 300 ×800 μm. The electron energy analyzer detected
photoelectrons perpendicular to the sample surface with an
acceptance angle of ±15. The spectra were recorded with
a step size of 0.1 eV and a pass energy of 10 eV, which
provided an overall energy resolution better than 0.5 eV. The
binding-energy scale of all XPS spectra was referenced to the
EF, which was set to a binding energy of 0 eV [15]. The ZrHx
thin films in this work were examined before and after Ar+
sputtering for 600 s using an Ar+incident angle of 20at 4 keV
rastered over an area of 2 ×2mm
2.
C. Structural model and density-functional theory calculations
The geometry relaxation was performed using the
Perdue-Burke-Ernzerhof (PBE) functional [16] including the
Grimme–van der Waals density-functional theory (DFT)-D2
scheme [17]. The first-principle calculations were carried out
using DFT implemented in the Vienna ab initio simulation
package (VASP)[18] with an exchange-potential functional
using the general gradient approximation (GGA). A hydrogen-
potential suited for short bonds (hydrogen projected aug-
mented wave (PAW) potential H_h) with an energy cutoff
of 1050 eV was used. For the self-consistent calculations,
the PAW [19] method was used with the PBE and the
Heyd-Scuseria-Ernzerhof functionals of the GGA. The HSE06
functional [20] is a hybrid functional including a linear
combination of short and long-range PBE exchange terms and
a short-range Hartree-Fock term that improves the formation
energies and band gaps and peak positions. The screening
parameter of HSE06 was 0.2˚
A1with a plane-wave cutoff
energy of 700 eV. A 29 ×29 ×29 kgrid was used for the
standard PBE functional for the structure relaxation and an
11 ×11 ×1kgrid was used for the HSE06 functional as
the computational cost is significantly higher for the HSE06
in comparison to the PBE functional. Using the HSE06
exchange-correlation functional, the energy positions of the
bands are shifted by 1.8 eV toward higher energy relative to
the EF, and are thus much more realistic, in comparison to the
corresponding energy positions, using the PBE functional. The
charge-transfer calculations between the different elements
were made using standard Bader analysis [21].
III. RESULTS AND DISCUSSION
Figure 2shows Zr 3dcore-level XPS spectra of the
ZrHxthin films in comparison to α-Zr metal. As observed,
there is a significant chemical shift toward higher binding
energies for the film grown with hydrogen in the plasma
compared to the pure metal film, which is a consequence of the
higher electronegativity of H (χ=2.2) in comparison to Zr
(χ=1.33) [22]. For x=1.16, the Zr 3d5/2peak position was
determined to 179.38 eV. This is in good agreement with the Zr
3d5/2binding energy of 179.375 reported for bulk δ-ZrH1.74
[23] as well as 179 eV determined for bulk ZrH1.9material
[24] and close to the 180-eV binding energy determined for
bulk ZrH1.64 [25] and ZrH1.9[7]. Furthermore, the position
of the 3d5/2peak is more or less identical to the binding
energy determined for the reactively sputtered ZrH1.64 film in
Ref. [14] with 179.44 eV. Note the small energy shift toward
lower binding energy for the less hydrogen-rich film with
x=1.16 (179.38 eV) that is consistent with a smaller charge
transfer from Zr to H in this film in comparison to the x=1.64
film (179.44 eV). In addition, for the ZrH1.16 film, there is a
high-energy shift of +0.51 eV while between x=0.15 and
0.30, the Zr 3dcore-level high-energy shift is much smaller
(+0.06 eV) and is an indication of smaller average charge
transfer from Zr toward H in these samples.
The metallic reference sample has the Zr 3d5/2and Zr
3d3/2peak positions at 178.87 and 181.28 eV (2.41-eV peak
splitting) in good agreement with literature values [26]. The Zr
3d5/2-3d3/2spin-orbit splitting is the same in the films.Prior
to sputtering, the spectra showed structures of ZrO2following
air exposure, but at the Zr 3dedge after sputtering there are
no features connected to oxygen.
195103-2
ELECTRONIC PROPERTIES AND BONDING IN ZrHx. . . PHYSICAL REVIEW B 96, 195103 (2017)
Intensity (arb. units)
184 182 180 178
Binding Energy (eV)
x
=0.00
x
=0.15
x
=0.30
x
=1.16
3d5/2
3d3/2
Zr 3d XPS
FIG. 2. Zr 3dcore-level XPS spectra of the ZrHxthin films in
comparison to α-Zr metal. The 3d5/2,3/2spin-orbit splitting (2.41 eV)
is indicated by the horizontal arrow.
To evaluate the asymmetric tails toward higher binding
energies in each spectrum, a Doniac-Sunjic function [27]
corresponding to the electron-hole pair excitations created at
the Fermi level to screen the core-hole potential was applied.
The Doniac-Sunjic profile is essentially a convolution between
a Lorentzian with the function 1/E(1 α), where Eis the
binding energy of each peak and αis a parameter known as
the singularity index that is related to the electron density of
states at the Fermi level [28]. The intensities of the tails are
related to the amount of metallicity in the system due to the
coupling of the core-hole with collective electron oscillations.
For α-Zr, the singularity index is large (α=0.15) and the
3d5/2/3d3/2branching ratio is higher (1.75) than the statistical
value of 1.67 (5/3), signifying high conductivity. For x=0.15
and 0.30, the singularity index and branching ratio remain
almost the same as for α-Zr, but for x=1.16, the tail becomes
significantly smaller with a singularity index of only α=0.04
with a branching ratio that is smaller (1.61) than the statistical
ratio (1.67). This shows that the number of bands crossing the
EFare reduced and thus the expected conductivity.
Figure 3shows Zr 4pshallow core-level XPS spectra of
the ZrHxthin films in comparison to α-Zr metal. As in the
case of the Zr 3dlevel, the largest chemical shift of +0.68 eV
from 27.26 eV in the metal to 27.94 eV is found for the film
with x=1.16, while a smaller shift of 0.1 eV occur for x=
0.30 with a similar shift in the composition range x=0.15.
The binding energy determined for the Zr 4p3/2peak of the
Intensity (arb. units)
40 35 30 25 20
Binding Energy (eV)
ZrHx4p-edge
x=1.16
x=0.30
x=0.15
x=0.00
Zr
4p
4p1/2
4p3/2
O 2s
FIG. 3. Zr 4pcore-level XPS spectra of the ZrHxthin films in
comparison to α-Zr metal. The 4p3/2,1/2spin-orbit splitting (1.3 eV)
is indicated by the vertical lines, from Ref. [26].
hydride with x=1.16 is slightly lower than the 28.1- and the
28.0-eV values reported in Refs. [9] and [12], respectively. For
both these studies, the spectra were recorded for oxidized bulk
samples, where the interaction of Zr atoms with O causes the
Zr 4ppeak to shift to a higher binding energy. Finally, the
4p3/2peak position of the metal reference of 27.26 eV is in
agreement with the reported value of 27.1 eV in Ref. [26].
The observed binding energies of the α-Zr spin-orbit split
4p3/2,1/2levels are 27.26 and 28.60 eV (1.3 eV splitting). This
is consistent with tabulated values of 27.1 and 28.5 eV [26].
At the shallow Zr 4plevel, we observe both a significant
broadening of the spectra and a high-energy shift in the
binding energy relative to metallic α-Zr. Only for the film
with x=0.30 some oxygen is left even after sputtering.
This was also observed by time-of-flight energy elastic recoil
detection analysis (ToF-ERDA) showing a slightly higher
oxygen content of 0.7% for this film [14] compared to 0.2%
for the films with x=0.15 and 1.16.
Figure 4shows a set of high-resolution VBs XPS spectra
of the ZrHxthin films (0.15, 0.30, and 1.16) in comparison
to α-Zr metal. In the region 0–2 eV from the EF, the spectra
are dominated by the Zr 4dvalence states with a 4d5/23/2
spin-orbit splitting of 1.3 eV, most clearly observed in
pure α-Zr (x=0). For x=0.15, the shape of the VB-XPS
spectrum is very similar to that of pure α-Zr with a larger
broadening and appears to be very little affected by the small
hydrogen content. On the contrary, for x=1.16, a major
spectral redistribution of intensity changes of the shape of
the Zr 4dband is further broadened so that the spin-orbit
splitting is no longer resolved. The most prominent new
feature appears around 6 eV binding energy and is due to the
195103-3
MAGNUSON, SCHMIDT, HULTMAN, AND HÖGBERG PHYSICAL REVIEW B 96, 195103 (2017)
Intensity (arb. units)
12 10 8 6 4 2 0 -2
Binding Energy (eV)
ZrHx Valence band
x=1.16
x=0.30
x=0.15
x=0.00
Zr
4d
Zr-H
hybr.
EF
4d3/2
4d5/2
O 2p
FIG. 4. Valence-band XPS spectra of the ZrHxthin films in
comparison to α-Zr metal.
strong H 1s–Zr 4dhybridization and bonding. For x=0.30, a
significant broadening of the Zr 4dvalence states at 1eVis
observed. The largest change is in the double-peaked feature
between 6 and 7.8 eV. This feature is consistent with previous
observations in bulk material [23] with δ-ZrH1.52,δ-ZrH1.65,
and δ-ZrH1.74, where the height of the Zr 4dpeak close to EF
decreases and the Zr–H bonding peak increases with increasing
hydrogen content. These observations suggested a donation
of Zr 4delectrons toward the Zr–H bond, while the Zr–Zr
bond weakens. For higher hydrogen content (x=1.94) the
states closest to the EFwere enhanced due to a fcc-fct phase
transition [9]. However, bulk materials are often hampered by
the appearance of superimposed O 2pstates between 5 and
10 eV, i.e., in the same energy region as the Zr–H bonding
peak. As ZrO2is known to have prominent peaks of O 2p
states between 5 and 10 eV and O 2sstates around 22.5 eV
from the EF[12]. Thus, in bulk materials, the contribution of
the Zr–H bond peak cannot be fully distinguished or separated
(disentangled) from the O 2pstates due to oxygen impurities.
Figure 5shows calculated density of states (DOS) for
crystalline δ-ZrHxfor x=0, 0.5, 1.0, 1.5, and 2.0. As
observed, for x=0, the states within 4 eV from the EFare
dominated by Zr 4dstates while strong Zr 4pstates occur at
29–30 eV from EF. This is also the case for increasing xbut a
prominent peak in the energy region 5–10 eV is mainly due to
H1sstates that does not exist in pure Zr. At 5–10 eV below EF,
bonding H–H s-,ZrHd-, and Zr–Zr d--interactions
occur while antibonding H–H s-interactions appear above
EF. We find that the H 1scharacter in δ-ZrHxstrongly depends
on the hydrogen content and increases with x. Notably, a deep
DOS minimum at 2–4 eV occurs for x=0.5 when hydrogen
is introduced. Closer to the EF,Zr4d-egstates dominate at
Density of States (DOS/eV/atom)
-35 -30 -25 -20 -15 -10 -5 0
Energy (eV)
Zr s
Zr p
Zr d
H s
H p
ZrH0.5
Zr
ZrH1.0
ZrH1.5
ZrH2.0
Zr 4p
FIG. 5. Calculated DOS for α-Zr and δ-ZrHx,wherex=0.5,
1.0, 1.5, and 2.0. To enhance the valence bands, the intense Zr 4p
states at 30 eV were scaled down by a factor of 5.
0–2 eV while H 1sstates are negligible. For δ-ZrH2, there is
an intense peak at the EFof Zr 4d-t2gcharacter that dominates
and causes instability of the crystal structure at 0 K. At finite
temperature, phonon interactions and tetragonal (Jahn-Teller)
distortion of the structure leads to a splitting of the intense
t2gpeak at the EFand a pseudogap develops. This leads to
lowering of the total energy of the system and restabilization
of the structure. Thus, the t2gstates at EFare reduced at
finite temperatures and likely reduce the conductivity and
other transport properties. This is consistent with previous
DFT studies including the δ- and εphases [2932]ofZrH
x
while comparisons to experimental electronic structure studies
of the Zr–H bonds of thin films have been lacking. Thus, the
order and bonding of the hydrogen atoms in the octahedral
sites has not been well understood.
Figure 6shows the structures (left) and calculated total
energies (right) of ZrHxfor x=0.5, 1.0, 1.5, and 2.0 (top
to bottom). Here, Jahn-Teller distortion is predicted for high
hydrogen content (x>1), where the cubic δstructure becomes
unstable and lowers the symmetry. On the contrary, for low
hydrogen content, the cubic structure is stable (x=0.5) or
metastable (x=1.0). For x=2.0, the minimum of c/a is
located at 0.886, for x=1.5 at 0.868, while for x=1 there is
a metastable state for c/a =1.0 and a lowest stable state for
c/a =0.792. Thus, the cubic δ-ZrHxphase is metastable as
thin film up to x=1.16, while for higher H contents the ZrHx
structure is predicted to be distorted by the Jahn-Teller effect.
195103-4
ELECTRONIC PROPERTIES AND BONDING IN ZrHx. . . PHYSICAL REVIEW B 96, 195103 (2017)
FIG. 6. (Left panel): Model structures. (Right panel): Calculated
total energies for ZrHx,wherex=0.5, 1.0, 1.5, and 2.0.
To evaluate the charge-transfer dependence on x,we
calculated the Bader charges that are listed in Table Itogether
with the lattice parameters for different compositions. Unit
cells of δ-ZrHx(Fig. 6, left panels), where the H atoms
were replaced by vacancies, were used as model systems for
different x. Among the hydrides, the structure and stability
properties of ZrH2have been of particular interest where the
attention has been focused on the δ-εphase transition induced
by stress, and the influence of the Jahn-Teller distortion to
stabilize the structure [30,33]. As the electronegativity of Zr
(1.33) is lower than for H (2.2), the formal oxidation states of
H and Zr in ZrH2are 1 and +2, respectively. Generally, the
predicted charge transfer from Zr increases with increasing
xand lattice parameter. In comparison to the pure Zr and H
TABLE I. Calculated lattice parameters and charge transfer for
ZrHxfor different x. Experimental values for α-Zr are given in
parenthesis [34].
System a-lattice parameter c-lattice parameter Zr H
α-Zr 3.2577 (3.2327) 5.1358 (5.1493) 0
ZrH0.54.6484 4.6607 +0.439e0.878e
ZrH1.04.7457 4.7457 +0.845e0.845e
ZrH1.54.7554 4.75547 +1.175e0783e
ZrH2.04.7811 4.78117 +1.491e0.745e
elements, the Bader charge on the Zr atoms in, e.g., ZrH2,
increases significantly by +1.49e, i.e., significant charge is
withdrawn from the 4d25s2valence orbitals of Zr. According
to the Bader analysis of ZrH2, the charge is transferred
to the lowest unoccupied orbital, of the hydrogen atoms
(1s11s1.75), where the Bader charge decreases by 0.75e
on each atom. A distortion of the cubic δ-ZrHxstructure
causes a slight reduction of the charge transfer.
The combined 3dand 4pcore-level and 4dvalence-band
studies show several interesting effects. For the hydride
system, we observe a significant reduction of the Zr 4dband
within 3 eV from the EFinfluencing the shallow 4pcore level
by a broadening and an asymmetric redistribution of intensity
toward higher binding energy. There is no trace of oxygen on
the samples, except for x=0.30, where small signs of O 2p
and O 2sintensity can be distinguished. For the hydrides, a
bonding band at 6 eV below the EFdoes not occur in pure Zr
and is related to strong H 1s–Zr 4dhybridization. This is also
accompanied by a chemical shift with binding energies that
are 0.6 eV higher than pure α-Zr indicating significant charge
transfer toward H, in particular for x=1.16. In our x-ray
absorption near-edge structure study [34], a chemical shift
toward higher energies in comparison to Zr metal was found
to be due to changes in the oxidation state that depend on
the structure and the formation of Zr–H bonding at sufficient
hydrogen loading. Previous valence-band XPS experiments
have shown a dominant Zr 4d peak at 1 eV and another peak
around 6.4 eV corresponding to a Zr–H bond [11], consistent
with DOS calculations [2931]. The chemical shift of the Zr
edges toward higher bonding energies confirms a significant
charge transfer from Zr toward H with increasing hydrogen
content. It has been argued that a tetragonal distortion of the
cubic δ-ZrHxstructure in combination with phonons make the
structure stable [30]. However, as shown in Fig. 6, this is only
the case for high H content, while at low H content, the cubic
structure is stable or metastable and does not distort. Thus, for
low H content, the cubic δstructure with tetrahedral Zr sites
favorably affects the materials properties in comparison to a
tentative octahedral coordination that for hydrogen is inhibited
according to the Hägg rules on size, coordination number, and
electronegativity.
For ZrHxfilms deposited by rDCMS, we have previously
shown that it is possible to grow films with tailored composi-
tions by altering the concentration of H2in the plasma [14].
The properties of the films resemble those of the parent Zr
metal where increasing hydrogen content in the films yields
higher resistivity seen from the values ranging from 70 to
120 μ cm in comparison to 42.6μ cm for α-Zr [35].
The measured hardness values are around 5.5 GPa, which
is harder than the 3 GPa determined for bulk α-Zr, using
nanoindentation [36].
In studies of bulk ZrHx, there have been problems with
oxide, as Zr is known to slowly form stable ZrO2on the
surface and at the grain boundaries [12]. Previous studies have
thus been limited to more or less oxidized polycrystalline
bulk materials that contain randomly ordered crystallites
with grain boundaries that affect the results of electronic
structure characterization. Moreover, investigations of the
electronic structures of understoichiometric transition-metal
hydrides have been scarce. Thin films synthesized by reactive
195103-5
MAGNUSON, SCHMIDT, HULTMAN, AND HÖGBERG PHYSICAL REVIEW B 96, 195103 (2017)
magnetron sputtering offer an alternative route to grow metal
hydrides such as ZrHxwith well-defined properties and at
a fraction of time (deposition time of about 2 min for an
800-nm film) compared to traditional hydration/percolation
processes that normally takes 1–2 days or weeks depending
on the H content [7]. A further advantage of the films
is the low level of contaminants, in particular, the oxygen
content that is of the range of 0.20.7% as determined
by ToF-ERDA [14]. Although 3dcore-level shifts in ZrHx
[14] have previously been observed, indicating charge-transfer
from Zr to H, a deeper analysis including branching ratio and
singularity index is necessary for the understanding how the
electronic properties influence the conductivity. For low H
content (x=0.15 and 0.30), the films contain a mixture of
metallic α-Zr and δ-ZrHxphases and the metallicity remains.
For higher H content (x=1.16), a homogeneous oriented,
textured, and understoichiometric δ-ZrHxphase is formed
with a lower singularity index and smaller branching ratio
that overall indicate more ceramic properties than for the
lower H contents. Producing highly stoichiometric and phase
pure materials is thus important for optimizing electrical and
hardness properties in a vast number of applications.
IV. CONCLUSIONS
Early transition-metal hydrides are a new class of thin-film
materials where the cubic δ-ZrHxphase is metastable up to
x=1.16, while for higher H contents the ZrHxstructure is
predicted to be distorted by the Jahn-Teller effect both in thin
films and bulk synthesized materials. By combining valence-
band and shallow core-level x-ray photoelectron spectroscopy
with electronic structure calculations, we have investigated the
electronic structure and chemical bonding of ZrHxthin films
with x=0.15, 0.30, and 1.16 in comparison to α-Zr metal. For
the understoichiometric δ-ZrH1.16 film, there are significant
chemical shifts of 0.68 and 0.51 eV for the 4p3/2and 3d5/2
peak positions toward higher energies compared to the metal
values of 27.26 and 178.87 eV. We find that even though there
is a significant charge transfer from the Zr 4dstates toward
the H 1sstates in δ-ZrH1.16, there are still states crossing the
Fermi level that signify metallicity. This is due to the fact
that there is a significant asymmetric redistribution of spectral
intensity from the shallow Zr 4pcore level toward the Zr 4d
valence band that accompanies the charge-transfer. There are
important changes at the Fermi level with a splitting of the 4d
band into a pseudogap where the Fermi level is located in a
local minimum that stabilize the structure and minimize the
total energy of the system. For the hydrides, a bonding band
around 6 eV below the Fermi level does not occur in pure Zr
and is related to strong H s–Zr 4dhybridization. This is also
accompanied by a chemical shift with binding energies that are
0.68 eV higher than pure α-Zr, indicating significant charge
transfer toward H, in particular for x=1.16.
ACKNOWLEDGMENTS
The research leading to these results has received funding
from the Swedish Government Strategic Research Area in
Materials Science on Functional Materials at Linköping Uni-
versity (Faculty Grant SFO-Mat-LiU No. 2009-00971). M.M.
acknowledges financial support from the Swedish Energy
Research (Grant No. 43606-1), the Swedish Foundation for
Strategic Research (SSF) (Grant No. RMA11-0029) through
the synergy grant FUNCASE, and the Carl Trygger Foundation
(Grants No. CTS16:303 and No. CTS14:310). Dr. Grzegorz
Greczynski is acknowledged for assistance with the measure-
ments of the Zr 3dcore-level XPS spectra.
[1] G. D. Moan and P. Rudling, in Zirconium in the Nuclear Indus-
try: Thirteenth International Symposium (ASTM International,
West Conshohocken, PA, 2002).
[2] M. P. Puls, The Effect of Hydrogen and Hydrides on the Integrity
of Zirconium Alloy Components (Springer, London, 2012).
[3] M.V.C.Sastri,MetalHydrides: Fundamentals and Applications
(Springer-Verlag, Berlin, 1998).
[4] L. Holliger, A. Legris, and R. Besson, Hexagonal-based ordered
phases in H-Zr, Phys. Rev. B 80,094111 (2009).
[5] JCPDS-International Centre for Diffraction Data. Zirconium
Hydride (ZrH2) Ref. ID 00-017-0314, http://www.icdd.com/.
[6] W. M. Mueller and J. P. Blackledge, Metal Hydrides (Academic,
New York, 1968).
[7] T. Riesterer, Electronic structure and bonding in metal hydrides,
studied with photoelectron spectroscopy, Z. Phys. B 66,441
(1987).
[8] R. C. Bowman, Jr., B. D. Craft, J. S. Cantrell, and E. L.
Venturini, Effects of thermal treatments on the lattice properties
and electronic structure of ZrHx,Phys.Rev.B31,5604 (1985).
[9] J. H. Weaver, D. J. Peterman, D. T. Peterson and A. Franciosi,
Electronic structure of metal hydrides. IV. TiHx,ZrH
x,HfH
x,
and the fcc-fct lattice distortion, Phys.Rev.B23,1692 (1981).
[10] H. A. Jahn and E. Teller, Stability of polyatomic molecules in
degenerate electronic states. I. Orbital degeneracy, Proc. R. Soc.
London A 161,220 (1937).
[11] M. Uno, K. Yamada, T. Maruyama, H. Muta, and S. Yamanaka,
Thermophysical properties of zirconium hydride and deuteride,
J. Alloys Compd. 366,101 (2004).
[12] B. W. Veal, D. J. Lam, and D. G. Westlake, X-ray photoemission
spectroscopy study of zirconium hydride, Phys. Rev. B 19,2856
(1979).
[13] T. B. Massalski, J. L. Murray, L. H. Bennett, and H. Baker,
Binary Alloy Phase Diagrams (American Society for Metals,
Metals Park, OH, 1986).
[14] H. Högberg, L. Tengdelius, M. Samuelsson, F. Eriksson, E.
Broitman, J. Lu, J. Jensen, and L. Hultman, Reactive sputtering
of δ-ZrH2thin films by high power impulse magnetron sputter-
ing and direct current magnetron sputtering, J. Vac. Sci. Technol.
A32,041510 (2014).
[15] S. Hüfner, Photoelectron Spectroscopy (Springer-Verlag, Berlin,
1996).
[16] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized
Gradient Approximation Made Simple, Phys.Rev.Lett.77,3865
(1996).
195103-6
ELECTRONIC PROPERTIES AND BONDING IN ZrHx. . . PHYSICAL REVIEW B 96, 195103 (2017)
[17] S. Grimme, Semiempirical GGA-type density functional con-
structed with a long-range dispersion correction, J. Comput.
Chem. 27,1787 (2006).
[18] G. Kresse and J. Furthmüller, Efficient iterative schemes for
ab initio total-energy calculations using a plane-wave basis set,
Phys. Rev. B 54,11169 (1996).
[19] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B
50,17953 (1994).
[20] J. Heyd, G. E. Scuseria, and M. Ernzerhof, Hybrid functionals
based on a screened Coulomb potential, J. Chem. Phys. 118,
8207 (2003).
[21] M. Yu and D. R. Trinkle, Accurate and efficient algorithm for
Bader charge integration, J. Chem. Phys. 134,064111 (2011).
[22] W. W. Por te rfie ld, Inorganic Chemistry: A Unified Approach,
2nd ed. (Academic, New York, 1993).
[23] S. Yamanaka, K. Yamada, K. Kurosaki, M. Uno, K. Takeda,
H. Anada, T. Matsuda, and S. Kabayashi, Characteristics of
zirconium hydride and deuteride, J. Alloys Compd. 99, 330
(2002).
[24] J. D. Corbett and H. S. Marek, A photoelectron spectroscopic
study of the zirconium monohalide hydrides and zirconium
dihydride. Classification of metal hydrides as conventional
compounds, Inorg. Chem. 22,3194 (1983).
[25] T. A. Sasaki and Y. Baba, Chemical-state studies of Zr and Nb
surfaces exposed to hydrogen ions, Phys. Rev. B 31,791 (1985).
[26] R. Nyholm and N. Mårtensson, Core level binding energies for
the elements Zr-Te (Z=40–52), J. Phys. C: Solid State Phys.
13,L279 (1980).
[27] S. Doniach and M. Sunjic, Many-electron singularity in X-ray
photoemission and X-ray line spectra from metals, J. Phys. C 3,
285 (1970).
[28] P. Nozières and C. T. De Dominicis, Singularities
in the x-ray absorption and emission of metals. III.
One-body theory exact solution, Phys. Rev. 178,1097
(1969).
[29] P. Zhang, B.-T. Wang, C.-H. He, and P. Zhang, First-principles
study of ground state properties of ZrH2original, Comput.
Mater. Sci. 50,3297 (2011).
[30] W. Wolf and P. Herzig, First-principles investigations of transi-
tion metal dihydrides, TH2:T=Sc, Ti, V, Y, Zr, Nb; energetics
and chemical bonding, J. Phys.: Condens. Matter 12,4535
(2000).
[31] T. Chihi, M. Fatmi, and A. Bouhemadou, Structural, mechanical
and electronic properties of transition metal hydrides MH2
(M=Ti,Zr,Hf,Sc,Y,La,VandCr),Solid State Sci.14,583
(2012).
[32] M. Gupta and J. P. Burger, Trends in the electron-phonon
coupling parameter in some metallic hydrides, Phys. Rev. B
24,7099 (1981).
[33] R. Quijano, R. de Coss, and D. J. Singh, Electronic structure
and energetics of the tetragonal distortion for TiH2, ZrH2,
and HfH2: A first-principles study, Phys.Rev.B80,184103
(2009).
[34] M. Magnuson, F. Eriksson, L. Hultman, and H. Högberg,
Bonding structures of ZrHxthin films by x-ray spectroscopy,
J. Phys. Chem. C (2017), doi:10.1021/acs.jpcc.7b03223.
[35] P. W. Bickel and T. G. Berlincourt, Electrical properties of
hydrides and deuterides of zirconium, Phys. Rev B 2,4807
(1970).
[36] M. Kuroda, D. Setoyama, M. Uno, and S. Yamanaka, Nanoin-
dentation studies of zirconium hydride, J. Alloys Compd. 368,
211 (2004).
195103-7
... The spectrum is de-convoluted into two zirconium doublets 3d 3/52 and 3d 5/2 which arise due to spin-orbit coupling having binding energies of 185.3 eV and 183.0 eV, respectively. The binding energy separation for these doublets turned out to be around 2.3 eV which is in close agreement with the reported value [17]. the phosphor [19] is responsible for the peak at 530.9 eV. ...
... the phosphor [19] is responsible for the peak at 530.9 eV. The bonding of Zr-O in the phosphor is ascribed to the existence of a peak at binding energy 530.4 eV [17]. This demonstrates that oxygen in the host lattice is bound by Na, Zr, and P. ...
Article
In this research article, Na(1-x)Zr2Smx(PO4)3 (x = 0.005-0.05) were synthesized via solution combustion route. The phase purity of the synthesized phosphors were explored by X-ray diffraction and Raman Spectroscopy. Rietveld refinement was used to compute the lattice parameters for Sm³⁺ doped NaZr2(PO4)3. The chemical states and morphological studies of the synthesized phosphors were carried out by X-ray photoelectron spectroscopy and field emission scanning electron microscope, respectively. A strong characteristic emission of Sm³⁺ ion was identified and the effect of doping concentration on PL spectra was extensively investigated. The bandgap was measured using diffuse reflectance spectra, and it was found in the range of 4.25 – 4.56 eV.
... For a-Zr [34], the main peak is dominated by the Zr-Zr paths at 3.160 Å (half diagonal, same notations as in Fig. 5) and 3.254 Å (cell edge), both with six nearest neighbors in the first coordination shell, which is close to our values obtained from θ/2θ XRD of the a-Zr bulk reference with 3.180 Å and 3.233 Å, respectively [37] [43] and Table III. Our XRD data is in excellent agreement with the bond lengths (3.179 Å and 3.232 Å) calculated from the JCPDS card [34]. ...
Preprint
The chemical bonding in an epitaxial ZrB2 film is investigated by Zr K-edge (1s) X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopies and compared to the ZrB2 compound target from which the film was synthesized as well as a bulk {\alpha}-Zr reference. Quantitative analysis of X-ray Photoelectron Spectroscopy spectra reveals at the surface: ~5% O in the epitaxial ZrB2 film, ~19% O in the ZrB2 compound target and ~22% O in the bulk {\alpha}-Zr reference after completed sputter cleaning. For the ZrB2 compound target, X-ray diffraction (XRD) shows weak but visible -111, 111, and 220 peaks from monoclinic ZrO2 together with peaks from ZrB2 and where the intensity distribution for the ZrB2 peaks show a randomly oriented target material. For the bulk {\alpha}-Zr reference no peaks from any crystalline oxide were visible in the diffractogram recorded from the 0001-oriented metal. The Zr K-edge absorption from the two ZrB2 samples demonstrate more pronounced oscillations for the epitaxial ZrB2 film than in the bulk ZrB2 attributed to the high atomic ordering within the columns of the film. The XANES exhibits no pre-peak due to lack of p-d hybridization in ZrB2, but with a chemical shift towards higher energy of 4 eV in the film and 6 eV for the bulk compared to {\alpha}-Zr (17.993 keV) from the charge-transfer from Zr to B. The 2 eV larger shift in bulk ZrB2 material suggests higher oxygen content than in the epitaxial film, which is supported by XPS. In EXAFS, the modelled cell-edge in ZrB2 is slightly smaller in the thin film (a=3.165 {\AA}, c=3.520 {\AA}) in comparison to the bulk target material (a=3.175 {\AA}, c=3.540 {\AA}) while in hexagonal closest-packed metal ({\alpha}-phase, a=3.254 {\AA}, c=5.147 {\AA}).
... Peaks derived from Ta were also detected at 17.0-31.0 eV, with O 2s 34) and Zr 4p 35,36) peaks also observed at binding energies close to those observed for Ta 4f. We assumed that the positions, shapes, and areas of the O 2s and 33,[37][38][39][40][41][42] . ...
Article
Full-text available
Metals that are used to reconstruct skeletal structures often interfere with magnetic resonance imaging (MRI) owing to differences in magnetic susceptibility; consequently, metals with lower magnetic susceptibilities need to be developed for use in implant devices. Herein, we investigated the corrosion properties of the Zr-14Nb-5Ta-1Mo alloy, which exhibits low magnetic susceptibility and excellent mechanical properties. The pitting potential of Zr-14Nb-5Ta-1Mo was higher than that of pure Zr. The passive current density of Zr-14Nb-5Ta-1Mo also higher than that of pure Zr, which is ascribable to slow reconstruction of the initial passive film associated with the presence of Nb and Ta. XPS revealed that the passive film is enriched with Nb and Ta. Therefore, while the Zr-14Nb-5Ta-1Mo alloy exhibited a high initial passive current density in simulated body fluid, it formed a stable passive film that suppressed localized corrosion. Zr-14Nb-5Ta-1Mo is therefore a prospective implant-material alloy candidate.
... Hf has ability to absorb thermal neutrons, generated due to an elastic collision of fast neutrons with H atoms, making hafnium hydride suitable and prolonged lifetime candidate in the fast reactors to be fabricated as control rods for fast neutron flux without producing Helium gas as in contrast with Boron Carbide (B 4 C) [6,7]. The binary hydrides of metals including Ti, Zr, Hf, Sc, and Y possesses greater stability and absorbs more hydrogen gas by implying low equilibrium hydrogen pressure, which makes hydrogen absorption irreversible even at room temperature [8][9][10]. Hafnium may favor greater hydrogen storage capacity, as it possesses larger affinity with hydrogen [11]. ...
Article
Metallic hydride clusters have greater importance due to its unique physicomechanical properties. For solid-state hydrogen storage, (HfH2)n clusters has been considered a promising candidate because of high hydrogen capacity, low cost and larger interacting affinity between atoms. The structural and electronic properties of (HfH2)n clusters are investigated by employing the density functional theory. From the DFT calculations, it is found that Hf atom occupies central position while H atoms tends to occupy at vertex spots. Through stability analysis, the calculated binding energy and second order energy difference of (HfH2)n clusters increases from (HfH2)5 through (HfH2)30. The charge density distribution and results of Bader analysis revealed ionic bonding character between Hf and H atoms and transfer of electrons is observed from Hf to H atoms. The orbital overlapping contribution of the interacting Hf and H atom is also performed.
... The spectrum is de-convoluted into two peaks of zirconium doublets 3d 5/2 and 3d 3/2 which arises due to spin orbit splitting and having binding energy at 181.4 and 183.7 eV, respectively. The binding energy separation of these doublets is 2.3 eV which is in close agreement with the reported value i.e. 2.4 eV [31]. Fig. 4 (d) shows the high resolution spectrum of Sm 3d 5/2 core level which is fitted with single broad peak centered at 1070.2 eV and in the sample Sm exists in 3+ valence chemical state [32]. ...
Article
In this work, Trivalent Sm (0.5-6 mol. %) doped mixed magnesium-zirconyl phosphor is synthesized by low temperature solution combustion method using urea as fuel. Phase formation, surface, and optical studies of the synthesized powder is characterized by X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), field emission scanning electron microscope (FESEM), photoluminescence (PL) spectroscopy, and ultraviolet visible (UV-Vis) spectroscopy. XRD result reveals the formation of mixed MgO-ZrO2 (cubic-tetragonal) phase. XPS technique is employed to examine the elemental composition of the synthesized material and results confirm the presence of Mg, Zr, O and Sm on the surface of the sample. The photoluminescence study shows that the synthesized phosphor gives four emission peaks at 567 nm, 602 nm, 651 nm, and 713 nm due to the electronic transitions of Sm³⁺ ion when excited under 403 nm. The photometric results reveal that the phosphor emits in the orange-red region of the color gamut and can be use as a warm light phosphor. The optical band gap of 4.74 eV is estimated from diffuse reflectance spectrum.
Article
Full-text available
Hydrogen's significance in contemporary society lies in its remarkable energy density, yet its integration into the worldwide energy grid presents a substantial challenge. Exposing materials to hydrogen environments leads to degradation of mechanical properties, damage, and failure. While the current approach for assessing hydrogen's impact on materials involves mainly multiscale modeling and mechanical testing, there exists a significant deficiency in detecting the intricate interactions between hydrogen and materials at the nanoatomic scales and under in situ conditions. This perspective review highlights the experimental endeavors aimed at bridging this gap, pointing toward the imminent need for new experimental techniques that can detect and map hydrogen in materials’ microstructures and their site‐specific dependencies.
Article
Full-text available
The extension of X-ray photoelectron spectroscopy (XPS) to measure layers and interfaces below the uppermost surface requires higher X-ray energies and electron energy analysers capable of measuring higher electron kinetic energies. This has been enabled at synchrotron radiation facilities and by using lab-based instruments which are now available with sufficient sensitivity for measurements to be performed on reasonable timescales. Here, we detail measurements on buried interfaces using a Ga Kα (9.25 keV) metal jet X-ray source and an EW4000 energy analyser (ScientaOmicron GmbH) in the Henry Royce Institute at the University of Manchester. Development of the technique has required the calculation of relative sensitivity factors (RSFs) to enable quantification analogous to Al Kα XPS, and here we provide further substantiation of the Ga Kα RSF library. Examples of buried interfaces include layers of memory and energy materials below top electrode layers, semiconductor heterostructures, ions implanted in graphite, oxide layers at metallic surfaces, and core-shell nanoparticles. The use of an angle-resolved mode enables depth profiling from the surface into the bulk, and is complemented with surface-sensitive XPS. Inelastic background modelling allows the extraction of information about buried layers at depths up to 20 times the photoelectron inelastic mean free path.
Article
It is widely accepted that the characteristics of hydrides in the Zr-based nuclear fuel cladding are critical factors determining the integrity of the spent fuel. Therefore, to mitigate any harm caused by them, we need to gain a deep understanding of the underlying mechanical behavior involving the precipitated hydrides. Using the molecular dynamics (MD) simulation method, we investigated the deformation mechanisms of embedded hydride in the polycrystalline zirconium matrix under uniaxial tensile loading. The plasticity of hydridized zirconium is controlled by the slip transmission from the metallic matrix into the embedded hydride, for which the considerable stress concentration at the metal-hydride interface needs to precede. The proposed mechanism manifests the size-dependent plasticity in the hydride, where the morphology of embedded hydride plays an important role in allowing the slip transmission through the matrix-hydride interface.
Article
The dehydrogenation temperature of LiAlH4 was significantly reduced by the production of mixtures with ZrCl4. Stoichiometric 4:1, and 5 mol % mixtures of LiAlH4 and ZrCl4 were produced by ball milling at room temperature and −196 °C, and tested for dehydrogenation at low temperature. Cryogenic ball-milling resulted in an effective way to produce reactive mixtures for hydrogen release; because of achieving small aggregates size (5–20 μm) in 10 min of cryomilling while preventing substantial decomposition during preparation. Dehydrogenation reaction in the mixtures LiAlH4/ZrCl4 started around 31–47 °C under different heating rates. Partial dehydrogenation was proved at 70 °C: 4.4 wt % for the 5 mol% ZrCl4–LiAlH4 mixture, and 3.4 wt % for the best 4:1 stoichiometric mixture. Complete dehydrogenation up to 250 °C released 6.4 wt% and 4.1 wt%, respectively. Dehydrogenation reactions are exothermic, and the LiAlH4/ZrCl4 mixtures are unstable and difficult to handle. The activation energy of the exothermic reactions was estimated as 113.5 ± 9.8 kJ/mol and 40.6 ± 6.6 kJ/mol for 4LiAlH4+ZrCl4 and 5%mol ZrCl4+LiAlH4 samples milled in cryogenic conditions, respectively. The dehydrogenation pathway was changed in the LiAlH4/ZrCl4 mixtures as compared to pure LiAlH4. Dehydrogenation reaction is proposed to form Al, LiCl, Zr, and H2 as main products. Modification of the dehydrogenation reaction of LiAlH4 was achieved at the cost of reducing the total hydrogen release capacity.
Poster
Full-text available
Circular giving details of the availability of a Stand Alone Report (SAR). This type of report, available from A.N.T. International, is equivalent to a book.
Article
Full-text available
Reactive sputtering by high power impulse magnetron sputtering (HiPIMS) and direct current magnetron sputtering (DCMS) of a Zr target in Ar/H2 plasmas was employed to deposit Zr-H films on Si(100) substrates, and with H content up to 61 at. % and O contents typically below 0.2 at. % as determined by elastic recoil detection analysis. X-ray photoelectron spectroscopy reveals a chemical shift of ∼0.7 eV to higher binding energies for the Zr-H films compared to pure Zr films, consistent with a charge transfer from Zr to H in a zirconium hydride. X-ray diffraction shows that the films are single-phase δ-ZrH2 (CaF2 type structure) at H content >∼55 at. % and pole figure measurements give a 111 preferred orientation for these films. Scanning electron microscopy cross-section images show a glasslike microstructure for the HiPIMS films, while the DCMS films are columnar. Nanoindentation yield hardness values of 5.5–7 GPa for the δ-ZrH2 films that is slightly harder than the ∼5 GPa determined for Zr films and with coefficients of friction in the range of 0.12–0.18 to compare with the range of 0.4–0.6 obtained for Zr films. Wear resistance testing show that phase-pure δ-ZrH2 films deposited by HiPIMS exhibit up to 50 times lower wear rate compared to those containing a secondary Zr phase. Four-point probe measurements give resistivity values in the range of ∼100–120 μΩ cm for the δ-ZrH2 films, which is slightly higher compared to Zr films with values in the range 70–80 μΩ cm.
Article
The variation in local atomic structure and chemical bonding of ZrHx (x=0.15, 0.30, 1.16) magnetron sputtered thin films are investigated by Zr K-edge (1s) X-ray absorption near-edge structure and extended X-ray absorption fine structure spectroscopies. A chemical shift of the Zr K-edge towards higher energy with increasing hydrogen content is observed due to charge-transfer and an ionic or polar covalent bonding component between the Zr 4d and the H 1s states with increasing valency for Zr. We find an increase in the Zr-Zr bond distance with increasing hydrogen content from 3.160 Å in the hexagonal closest-packed metal (α-phase) to 3.395 Å in the understoichiometric -ZrHx film (CaF2-type structure) with x=1.16 that largely resembles that of bulk δ-ZrH2. For yet lower hydrogen contents, the structures are mixed α and δ-phases, while sufficient hydrogen loading (x>1) yields a pure δ-phase that is understoichiometric, but thermodynamically stable. The change in the hydrogen content and strain is discussed in relation to the corresponding change of bond lengths, hybridizations, and trends in electrical resistivity.
Article
The electronic structures of TiHx, ZrHx, and HfHx have been studied using photoelectron spectroscopy and synchrotron radiation. Structures in the metal d-derived band within ~3 eV of the Fermi level EF and in the bonding band (~3-10 eV below EF) are compared with theory. In each dihydride, the bonding band center falls at -5.5 eV, at approximately the same energy as previously observed for the dihydrides of Sc and Y. Changes in the emission features near EF and at -7 eV have been observed in samples bridging the fcc-->fct distortion in ZrHx, 1.63<=x<=1.94. The changes at EF demonstrate the Jahn-Teller effect for the electronic states of ZrHx. The binding energies of the Ti 3p, Zr 4p, Hf 5p, and Hf 4f cores are observed to be greater than in the elemental metals, consistent with charge transfer to the hydrogen site.
Article
First-principles calculations have been carried out to investigate the structural, mechanic and electronic of transition metal hydrides MH2 (M = Ti, Zr, Hf, Sc, Y, La, V and Cr). It is found that TiH2 is mechanically unstable because of a negative C44 = −21.31 GPa and C11–C12 < 0, the same behavior can be found in MH2 (M = Zr, Hf, and Y) compounds. Also there is a strong interaction between M (Ti, Zr, Hf, Sc, Y, La, V and Cr) and H. On the other hand, the H–H bond orders are always negative or nil reason of brittleness.
Article
It is pointed out that the spectra of X-ray induced fast photoelectrons from metal should have a characteristic skew line shape resulting from Kondo-like many-electron interactions of the metallic conduction electrons with the accompanying deep hole in the final state. The same line shape should also occur for the discrete line spectra of X-rays emitted from metals. This mechanism could account for the well-known asymmetries observed for K alpha lines.
Article
Structural, mechanical, electronic, and thermodynamic properties of fluorite and tetragonal phases of ZrH2 are systematically studied by employing the density functional theory within generalized gradient approximation. The existence of the bistable structure for ZrH2 is mainly due to the tetragonal distortions. And our calculated lattice constants for the stable face-centered tetragonal (fct) phase with c/a=0.885 are consistent well with experiments. Through calculating elastic constants, the mechanically unstable characters of face-centered cubic (fcc) phase and fct structure with c/a=1.111 are predicted. As for fct0.885 structure, our calculated elastic constants explicitly indicate that it is mechanically stable. Elastic moduli, Poisson’s ratio, and Debye temperature are derived from elastic constants. After analyzing total and partial densities of states and valence electron charge distribution, we conclude that the Zr–H bonds in ZrH2 exhibit weak covalent feature. But the ionic property is evident with about 1.5 electrons transferring from each Zr atom to H. Phonon spectrum results indicate that fct0.885 structure is dynamically stable, while the fcc and fct1.111 structures are unstable in accord with the mechanical stability analysis.