ArticlePDF Available

Abstract and Figures

This letter reports on the pressure dependence of the optical absorption edge of ZnO in the rock-salt phase, up to 20 GPa. Both vapor-phase monocrystals and pulsed-laser-deposition thin films on mica have been investigated. Rock-salt ZnO is shown to be an indirect semiconductor with a band gap of 2.45+/-0.15 eV, whose pressure coefficient is very small. At higher photon energies, a direct transition is observed (4.6 eV at 10 GPa), with a positive pressure coefficient (around 40+/-3 meV/GPa between 5 and 19 GPa). These results are interpreted on the basis of first-principles electronic band structure calculations.
Content may be subject to copyright.
Optical properties and electronic structure of rock-salt ZnO under pressure
A. Segura
a)
and J. A. Sans
Institut de Cie
`
ncia dels Materials-Dpto. de
´
sica Aplicada, Universitat de Vale
`
ncia, Ed. Investigacio
´
,
E-46100 Burjassot (Vale
`
ncia), Spain
F. J. Manjo
´
n
Departimento de
´
sica Aplicada, Universitat Polite
`
cnica de Vale
`
ncia, 03801 Alcoy (Alicante), Spain
A. Mun
˜
oz and M. J. Herrera-Cabrera
Departimento De
´
sica Fundamental II, Universidad de la Laguna, 38204 La Laguna (Tenerife), Spain
Received 10 February 2003; accepted 14 May 2003
This letter reports on the pressure dependence of the optical absorption edge of ZnO in the rock-salt
phase, up to 20 GPa. Both vapor-phase monocrystals and pulsed-laser-deposition thin films on mica
have been investigated. Rock-salt ZnO is shown to be an indirect semiconductor with a band gap of
2.45 0.15 eV, whose pressure coefficient is very small. At higher photon energies, a direct
transition is observed 4.6 eV at 10 GPa, with a positive pressure coefficient around 40
3 meV/GPa between 5 and 19 GPa. These results are interpreted on the basis of first-principles
electronic band structure calculations. © 2003 American Institute of Physics.
DOI: 10.1063/1.1591995
Zinc oxide ZnOis attracting a renewed attention owing
to its potential applications in ultraviolet optoelectronic
devices.
1
At room pressure, ZnO crystallizes in the wurtzite
structure W-ZnOand transits to the rock-salt structure RS-
ZnO at about 9 GPa.
2
Upon decompression, the reverse
transition in bulk crystals has a large hysteresis and occurs
below 4 GPa.
3,4
Several authors have claimed to be able to
obtain metastable RS-ZnO at ambient conditions.
5,6
This
possibility has recently been proved for nanocrystalline
samples in which the RS phase is maintained at ambient
conditions after a pressure cycle up to 16 GPa.
7
It has also
been proved that thin films of Mg
x
Zn
1x
O grown by pulsed
laser deposition PLDhave the RS structure for x 0.5 with
band-gap energies larger than 5 eV and increasing with the
Mg content.
8
Therefore, the knowledge of the electronic
structure of RS-ZnO and, more specifically, of the nature and
value of its band-gap, is a subject of interest for future tech-
nological applications of these nanocrystallites or thin films.
Apart from being a powerful material preparation tech-
nique, high pressure is also a very efficient tool for under-
standing the electronic structure of semiconductors.
9
How-
ever, relatively few experimental results have been reported
on the electronic structure of W-ZnO under pressure,
10–12
as
compared to other II-VI compounds. The same holds for
RS-ZnO, whose electronic properties have been investigated
only in theoretical papers,
13–15
which predict an indirect
semiconductor character.
In this letter, we report on the pressure dependence of
the absorption edge of RS-ZnO measured at room tempera-
ture RT and discuss the results on the basis of first-
principles density-functional-theory calculations.
For the optical absorption measurements, both W-ZnO
bulk single crystals and thin films were used. Large bulk
single crystals were grown by the vapor-phase method
16
and
broken into small splints with parallel faces and 15-
m thick
for measurements in the diamond anvil cell DAC. Thin
films were prepared by PLD on mica monocrystalline sub-
strates.
The target for the thin-film preparation was a com-
pressed pellet of 5N ZnO powder annealed at 950°C for 6 h
in air atmosphere. Samples were prepared at relatively low
temperature (400 °C) with a dynamically controlled atmo-
sphere of 5N oxygen at 210
4
mbar and then subjected to
6 h annealing in air at higher temperatures up to 600°C).
For optical absorption measurements in the UV-Vis-NIR
range under pressure, a sample was placed together with a
ruby chip into a 200-
m-diameter hole drilled on a 50-
m-
thick Inconel™ gasket and inserted between the diamonds of
a membrane-type DAC.
17
Methanolethanolwater 16:3:1
was used as a pressure transmitting medium, and the pressure
was determined through the ruby luminescence linear
scale.
18
The optical setup was similar to the one described in
Ref. 17. It consists of a Xe lamp, fused silica lenses, reflect-
ing optics objectives, and an UV-Vis spectrometer, which
allows for transmission measurements up to the absorption
edge of IIA diamonds about 5.5 eV.
The electronic structure at different pressures of RS-ZnO
has been calculated through the ab initio total energy
pseudopotential plane wave method using the density func-
tional theory DFT in the framework of the local density
approximation LDA. The semi-core 3d electrons of Zn are
treated as forming part of the valence states. Plane waves up
to 130 Ry energy cutoff were used in order to have highly
convergent results.
As regards the wurtzite phase, in both the bulk and thin
film samples, a monotonous blueshift of the absorption edge
is observed as pressure increases. In the experiment with
bulk ZnO samples, the use of a 15-
m-thick sample allows
only for the observation of the low-energy tail of the absorp-
tion edge in W-ZnO. The transition from the W-ZnO bulk
sample to the NaCl phase occurs at about 9.5 0.2 GPa, and
is observed as a neat change in the shape of the absorption
coefficient see Fig. 1b兲兴. The absorption edge exhibits a
structure related to the wurtzite phase up to 16 GPa. Above
a
Electronic mail: alfredo.segura@uv.es
APPLIED PHYSICS LETTERS VOLUME 83, NUMBER 2 14 JULY 2003
2780003-6951/2003/83(2)/278/3/$20.00 © 2003 American Institute of Physics
Downloaded 23 Jul 2003 to 158.42.129.75. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/aplo/aplcr.jsp
that pressure, and contrary to what happens for most high-
pressure phases of II-VI semiconductors, the sample does not
exhibit any trace of light scattering and the absorption edge
can be accurately measured. Figure 2 shows the absorption
edge in RS-ZnO as measured in the down-stroke from 19 to
4 GPa with the bulk sample.
The low-energy tail of the absorption edge has a qua-
dratic dependence on the photon energy at all pressures, as
illustrated in the inset of Fig. 2 for P14.5 GPa. Then this
tail can be assigned to an indirect transition with a band-gap
of 2.47 0.02 eV at 14.5 GPa. At other pressures from 4.7
to 19.9 GPa the band-gap values vary between 2.33 and
2.61 eV, but do not show any clear trend of pressure depen-
dence. The average value would be 2.45 0.15 eV. This is
about half the theoretical values predicted by correlated
HartreeFock calculations 5.54 eV
13
or by quasiparticle
methods 4.51 eV.
15
Our result is also to be compared to the extrapolated
value of the band gap of RS-Mg
x
Zn
1 x
O thin films for x
0, that is slightly higher than 3 eV.
8
This is a normal dis-
agreement, as transmission measurements in thin films are
not sensitive to the low values of the absorption coefficient
typical of indirect transitions and, consequently, yield over-
estimated band-gap values.
Contrary to the low-energy tail, the absorption spectrum
at higher photon energies exhibits a clear pressure depen-
dence. Above 3.5 eV see Fig. 2, the absorption edge be-
comes steeper and blueshifts with a increasing pressure. The
pressure coefficient, as determined from the shift of the pho-
ton energy at a constant absorption coefficient, is about 34
2 meV/GPa between 5 and 19 GPa.
In ZnO/mica thin films the phase transition to the NaCl
phase is observed also at about 9.5 0.2 GPa, as a sudden
change in the absorption edge, from the step-like shape typi-
cal of the wurtzite phase, with its exciton-related maximum,
to a structureless absorption tail see Fig. 1a兲兴. As discussed
earlier, the small thickness of the films prevents the observa-
tion of the indirect transition at 2.47 eV. Instead, an absorp-
tion edge at photon energies higher than 4.5 eV is observed,
which is likely related to an allowed direct transition. Figure
3 shows this absorption edge at several pressures. The inset
of Fig. 3 shows the pressure dependence of the direct gap, as
determined from extrapolating the linear part of the
2
ver-
sus h
v
plot at each pressure. The pressure coefficient of this
absorption edge is 40 3 meV/GPa.
Figure 4 shows the pressure dependence of the ab initio
calculated electronic band structure, on the basis of which we
will discuss the experimental results. RS-ZnO turns out to be
an indirect semiconductor. Its conduction band minimum
CBM, located at the point of the Brillouin zone BZ,is
about 1.1 eV above the valence band maxima VBM, lo-
cated at the L point and midway in the –K or ⌺兲 direction.
The fundamental transition is correctly predicted to be indi-
rect, even if the band gap is underestimated 1.1 eV versus
the experimental value of 2.45 eV.Inthe point of the BZ
in the RS structure (O
h
point group p and d states belong to
different representations and do not mix. Away from the
point, p and d states mix and the resulting pd repulsion
results in the upwards dispersion of p bands in directions
–K 共⌺兲 and –L 共⌳兲 of the BZ Consequently, the VBM
FIG. 1. Change of the absorption coefficient at the wurtzite-to-rock-salt
transition in ZnO thin films on mica a and bulk samples b.
FIG. 2. RT absorption edge of rock-salt ZnO at different pressures, as mea-
sured in a bulk sample. Inset: square root of the absorption coefficient as a
function of photon energy at 14.5 GPa.
FIG. 3. RT absorption edge of rock-salt ZnO at different pressures, as mea-
sured in a thin film deposited on mica. Inset: pressure dependence of the
photon energy at a constant absorption coefficient (
10
5
cm
1
).
279Appl. Phys. Lett., Vol. 83, No. 2, 14 July 2003 Segura
et al.
Downloaded 23 Jul 2003 to 158.42.129.75. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/aplo/aplcr.jsp
occurs away from the point and the material has indirect
character.
The calculated pressure coefficient of the indirect band
gaps L– and ⌫兲 are both positive (25 meV/GPa).
This result contrasts with the pressure insensitivity of the
observed indirect transition and prevents an unambiguous
assignment. On the opposite, the direct transition at higher
photon energy can be reasonably assigned. Direct transitions
at X and L points have very large energies and pressure
coefficients see Fig. 4, then the most reasonable assignment
seems to be the direct transition. This assignment is
supported by the comparison of the theoretical pressure co-
efficient for this direct transition 45.4 meV/GPa and the
experimental one (40 3 meV/GPa).
Table I compares the pressure coefficients and deforma-
tion potentials of the direct transition in ZnO wurtzite
and RS phases. The fact that both parameters are much larger
in the RS phase is a consequence of role played by pd
interaction in each phase. Pd repulsion in the valence band
has been proposed to be responsible for the so-called band-
gap anomaly in wurtzite II-VI
20
and in I-III-VI
semiconductors.
21
It has also been invoked as responsible for
the low deformation potential of the band gap in wurtzite and
zinc-blende II-VI materials.
22
According to that model, when
pd mixing is symmetry forbidden, one should expect de-
formation potentials for direct transitions at the point to be
close to those in light III-V zinc-blende compounds,
22
as it
actually happens in RS-ZnO. In the same direction, it is also
relevant to notice that the ‘band-gap anomaly,’
20,21
illus-
trated in the II-VI wurtzite semiconductors by the larger
band gap of ZnS 3.7 eV with respect to ZnO 3.35 eV,
does not occur in the RS phase: the band gap of RS-ZnO
2.45 eV is larger than the one of RS-ZnS 2.0 eV.
23
In summary, we have investigated the electronic band
structure of rock-salt ZnO in bulk crystals and thin films
observing the same wurtzite-to-rock-salt phase transition
pressure (9.5 0.2 GPa) in both types of samples. RS-ZnO is
an indirect gap semiconductor with a band gap of 2.45
0.15 eV at 13.5 GPa. An intense direct transition at
higher energy about 4.5 eV at 10 GPa, with a large positive
pressure coefficient (40 3 meV/GPa), has also been ob-
served and assigned to the lowest direct transition at the
point of the BZ. Further investigations are needed to solve
the discrepancy between the measured and calculated ener-
gies and pressure coefficients of the indirect band gaps.
The authors thank R. Lauck for kindly providing the
bulk single crystals. This work was supported through Span-
ish Government MCYT grants MAT2002-04539-C02-0102
and BFM2001-3309-C02-01 02. One of the authors
M.J.H.C. wish to thank the Spanish MCYT FPI fellowship
program.
1
H. Ohta, K. Kawamura, M. Orita, N. Sarukura, M. Hirano, and H.
Hosono, Electron. Lett. 36,9842000.
2
C. H. Bates, W. B. White, and R. Roy, Science 137,9931962.
3
F. J. Manjo
´
n, K. Syassen, and R. Lauck, High Press. Res. 22,2992002.
4
H. Karzel, W. Potzel, M. Ko
¨
fferlein, W. Scchiessel, M. Steiner, U. Hiller,
G. M. Kalvius, D. W. Mitchel, T. P. Das, P. Blaha, K. Schwarz, and M. P.
Pasternak, Phys. Rev. B 53, 11425 1996.
5
L. Gerward and J. Staun Olsen, J. Synchrotron Radiat. 2,2331995.
6
J. M. Recio, M. A. Blanco, V. Luan
˜
a, R. Pandey, L. Gerward, and J. Staun
Olsen, Phys. Rev. B 58, 8949 1998.
7
F. Decremps, J. Pellicer-Porres, F. Datchi, J. P. Itie
´
, A. Polian, F. Baudelet,
and J. Z. Jiang, Appl. Phys. Lett. 81,48202002, and references therein.
8
S. Choopun, R. D. Vispute, W. Yang, R. P. Sharma, T. Venkatesan, and H.
Shen, Appl. Phys. Lett. 80, 1529 2002.
9
W. Paul, High Pressure in Semiconductor Physics I, edited by T. Suski and
W. Paul, Semiconductors and Semimetals Vol. 54 Academic, San Diego,
1998,p.1.
10
R. L. Knell and D. W. Langer, Phys. Lett. 21,2701966.
11
D. R. Huffman, L. A. Schwalbe, and D. Schiferl, Solid State Commun. 44,
521 1982.
12
A. Mang, K. Reimann, and St. Ru
¨
benache, Solid State Commun. 94,251
1995.
13
J. E. Jaffe, R. Pandey, and A. B. Kunz, Phys. Rev. B 43, 14030 1991.
14
J. E. Jaffe, J. A. Snyder, Z. Lin, and A. C. Hess, Phys. Rev. B 62,1660
2000.
15
H. Q. Ni, Y. F. Lu, and Z. M. Ren, J. Appl. Phys. 91, 1339 2002.
16
R. Helbig, J. Cryst. Growth 15,251972.
17
R. Le Toullec, J. P. Pinceaux, and P. Loubeyre, High Press. Res. 1,77
1988.
18
G. J. Piermarini, S. Block, J. D. Barnett, and R. A. Forman, J. Appl. Phys.
46, 2774 1975.
19
S. Desgreniers, Phys. Rev. B 58, 14102 1998.
20
S. H. Wei and A. Zunger, Phys. Rev. B 37, 8958 1988.
21
J. E. Jaffe and A. Zunger, Phys. Rev. B 29, 1882 1984.
22
S. H. Wei and A. Zunger, Phys. Rev. B 60, 5404 1999.
23
S. Ves, U. Schwarz, N. E. Christensen, K. Syassen, and M. Cardona, Phys.
Rev. B 42,91131990.
FIG. 4. Electronic band structure of rock-salt ZnO along high-symmetry
directions of the BZ at several pressures, as calculated through ab initio
DFT-LDA pseudopotential method.
TABLE I. Pressure coefficient and deformation potential of the direct
transition in ZnO crystalline phases.
ZnO
phase
dE
gd
⌫⌫
/dP
meV/GPa
B
0
GPa
dE
gd
⌫⌫
/d ln V
eV
Wurtzite 24.5 2
a
142.6 2
b
3.50.4
RS 40 3
a
202.5 2
b
8.10.5
a
This work.
b
Reference 19.
280 Appl. Phys. Lett., Vol. 83, No. 2, 14 July 2003 Segura
et al.
Downloaded 23 Jul 2003 to 158.42.129.75. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/aplo/aplcr.jsp
... The band structures of seven ZnO types are plotted along the high symmetry lines of the Brillouin zone by the LDA and LDA + U, as shown in Figs. 3 and 4. The calculated band gaps for each type are shown in Table 3 (Zagorac et al. 2014;Cui et al. 2009;Amrani et al. 2006;Usuda et al. 2002;Mang et al. 1995;Schroer et al. 1993;Oshikiri and Aryasetiawan Page 5 of 23 1177 2000; Xu and Ching 1993;Segura et al. 2003;Kaczkowski 2012;Sponza et al. 2015;Bashyal et al. 2018). LDA method, for the B4 phase, a direct band gap at the G point with a value of 0.796 eV, several in agreement with the theoretical results (Cui et al. 2009;Amrani et al. 2006;Usuda et al. 2002), but this value is much smaller than the experimental value of 3.44 eV (Mang et al. 1995). ...
... LDA method, for the B4 phase, a direct band gap at the G point with a value of 0.796 eV, several in agreement with the theoretical results (Cui et al. 2009;Amrani et al. 2006;Usuda et al. 2002), but this value is much smaller than the experimental value of 3.44 eV (Mang et al. 1995). For the B1 phase, an indirect band gap at the (L-G) of 1.174 eV, this value is similar to the theoretical data, but is smaller than the experimental data of 2.45 eV (Segura et al. 2003). The large difference between our calculated and experimental values may due to the using of LDA, which overestimates the energy localization value of Zn-3d (Powell et al. 1971), thus overestimating the hybridization between Zn-3d and O-2p states, resulting in an increased valence band bandwidth and low band gap. ...
Article
Full-text available
We have used first-principles density-functional theory calculations to determine the structural, electronic and optical properties of four known ZnO phases (B1, \(Fm\overline{3} m\)), (B2, \(Pm\overline{3} m\)), (B3, \(F\overline{4} 3m\)), (B4, \(P6_{3}mc\)) and three predicted phases of ZnO modification (GeP, I4mm), (5–5, \(P6_{3}/mmc\)), (NiAs, \(P6_{3}/mmc\)). The difference of data calculated by LDA and LDA + U has been compared. In addition, the calculated ground state energies and structural parameters of these phases are in consistent with the reported theoretical and experimental results. We have discussed the band structures, densities of states and chemical bonding of the seven types. Further, we have compared the calculated values of the dielectric function of the B4 phase with the experimental values. In addition, the optical constants of the seven types of ZnO have been calculated by LDA + U and LDA, respectively, the results obtained by the two methods are compared and analyzed.
... As it was proved by Grinberg et al., [30] the energy of the conduction band increases with applied pressure in some host materials. [31][32][33][34] Therefore the 5d1 state of Ce(1) arises from the conduction band leading to the gradual enhancement of its contribution in the emission spectra. As can be clearly seen in the emission spectra recorded at ambient pressure, the emission spectrum is dominated by the band associated with electron transitions between the energy levels of the Ce(2) ion. ...
Article
Full-text available
For overcoming the unsatisfactory sensitivity of optical manometers and the susceptibility of manometric parameters to temperature changes, a series of Ce³⁺‐doped La6Sr4(SiO4)6F2 phosphors with multicolor emissions are developed. Excited by 375 nm, the resultant compounds exhibit asymmetric and broad emission bands, which can be easily deconvoluted into two peaks originating from seven to ninefold coordinated Ce³⁺ sites, where their intensities are dependent on the excitation wavelength. Through analyzing the pressure‐related luminescence intensity ratio (LIR) derived from these two Ce³⁺ sites, it is clear that the final manometer exhibits an unprecedentedly high relative sensitivity of ≈ 425% GPa⁻¹ at 0.76 GPa (the highest value reported to date). As pressure increases, the emitting color tunes from yellow to blue, resulting in the possibility of visual pressure sensing. Furthermore, based on the color coordinates of the phosphor material, its pressure sensing capability is discussed, in which the maximum sensitivities are 24.38 and 27.74% GPa⁻¹ at 0.76 GPa, when x‐coordinate and y‐coordinate are employed, respectively. Additionally, the LIR value is only slightly affected by temperature, leading to a thermal invariability manometric factor (TIMF) > 400 K GPa⁻¹. The findings indicate that Ce³⁺‐doped La6Sr4(SiO4)6F2 phosphors are good luminescent candidates for optical and visual multi‐mode manometers.
... where ¢ 0 = 191 GPa and ¢ 0 A = 3.54 for ZnO with the rocksalt structure reported by Decremps et al., 2) and © (= µ/µ 0 ) is the ratio of the densities under applied and ambient pressures, respectively. This ratio can be equated as 3 where d 0 and d are the lattice spacings at the corresponding conditions, and further as © = E/E 0 where E and E 0 are the peak energies under applied and ambient pressures, respectively. This equality arises using the relations as µ ³ m/d 3 and E ³ 1/2d sin ª where m is the mass per unit volume and ª is the X-ray reflection angle. ...
Article
Full-text available
Zinc oxide (ZnO) with a rocksalt crystal structure is attractive because of the bandgap which lies in the range of visible light absorption (1.22.6 eV). However, the rocksalt structure is not stable at ambient pressure and temperature according to an equilibrium phase diagram. Nevertheless, this study demonstrates, for the first time, that it is possible to realize a 100% fraction of the rocksalt structure at ambient pressure and temperature. ZnO powder is initially processed by severe plastic deformation under high pressure through a technique of high-pressure torsion (HPT). The HPT-processed ZnO is then examined using a high-pressure application system available at BL04B1 of SPring-8 and in situ X-ray diffraction (XRD) analysis is conducted under high pressures at elevated temperatures. It is shown that the initial presence of the rocksalt structure produced by the HPT process is effective to attain a 100% fraction of the rocksalt structure. [
... One would expect E g of MZ20 and MZ30 to be approximately 3.51 and 4.05 eV, respectively, assuming the previously reported E g = 2.45 eV for pure rs-ZnO [35], E g = 7.77 eV for MgO [36] and b = 0 in Equation (3). However, there is almost no difference between the observed band gaps of pure rs-ZnO (2.71 eV), MZ20 (2.57 ...
Article
Full-text available
The possibility of doping ZnO in its metastable rock salt structure with Li, Na, and K intended to act as acceptor dopants was investigated. For the first time, MgxZn1−xO alloys and pure ZnO with a rock salt structure doped with Li, Na, and K metals was obtained by high-pressure synthesis from pure oxides with the addition of carbonates or acetates of the corresponding metals as dopant sources. Successful stabilization of the metastable rock salt structure and phase purity were confirmed by X-ray diffraction. Transmission electron microscopy was used to study the particle size of nanocrystalline precursors, while the presence of Li, Na, and K metals in rock salt ZnO was detected by electron energy-loss spectroscopy and X-ray photoelectron spectroscopy in MgxZn1−xO alloys. Electron paramagnetic resonance measurements revealed the acceptor behavior of Li, Na, and K dopants based on the influence of the latter on native defects and natural impurities in ZnO-MgO alloys. In addition, diffuse reflectance spectroscopy was used to derive band gaps of quenched rock salt ZnO and its alloys with MgO.
Article
Full-text available
Ultrafine-grained and heterostructured materials are currently of high interest due to their superior mechanical and functional properties. Severe plastic deformation (SPD) is one of the most effective methods to produce such materials with unique microstructure-property relationships. In this review paper, after summarizing the recent progress in developing various SPD methods for processing bulk, surface and powder of materials, the main structural and microstructural features of SPD-processed materials are explained including lattice defects, grain boundaries and phase transformations. The properties and potential applications of SPD-processed materials are then reviewed in detail including tensile properties, creep, superplasticity, hydrogen embrittlement resistance, electrical conductivity, magnetic properties, optical properties, solar energy harvesting, photocatalysis, electrocatalysis, hydrolysis, hydrogen storage, hydrogen production, CO2 conversion, corrosion resistance and biocompatibility. It is shown that achieving such properties is not currently limited to pure metals and conventional metallic alloys, and a wide range of materials are processed by SPD, including high-entropy alloys, glasses, semiconductors, ceramics and polymers. It is particularly emphasized that SPD has moved from a simple metal processing tool to a powerful means for the discovery and synthesis of new superfunctional metallic and nonmetallic materials. The article ends by declaring that the borders of SPD have been extended from materials science and it has become an interdisciplinary tool to address scientific questions such as the mechanism of geological and astronomical phenomena and the origin of life.
Article
We report the photoelectric properties and structural changes of ZnO nanoparticles (∼63 nm) under pressures of up to 22.9(5) GPa using in situ Raman spectroscopy, photocurrent measurements, and theoretical calculations. The ZnO nanoparticles show enhanced photocurrent (2.8 mA) and responsivity (4.78 × 10 6 mA W −1) under 365 nm irradiation in the wurtzite (B4) phase. On the contrary, the rock-salt (B1) phase, emerging post the wurtzite phase, exhibits a smaller band gap and decreased absorption coefficient, leading to reduced photocurrents and responsivity. These findings highlight the potential of high-pressure modulation to optimize the photoelectronic properties of ZnO nanoparticles for electronic device applications.
Article
The thickness-dependent atomic structures of two-dimensional (2D) few-layer (FL) ZnO are systematically investigated by the first-principles calculations. It is found that the structural transformation between thinner FL ZnO with graphitic structure (FL gZnO) and thicker FL ZnO with wurtzite structure (FL wZnO) takes place at the critical thickness of 9–12 Zn-O atomic layers. At the thickness of 9–12 layers, both graphitic and wurtzite structures can coexist at room temperature. In FL gZnO, the interlayer interaction is a long-range Coulomb interaction, and the charge population of Zn and O inside does not change during the structural transformation. Moreover, we demonstrate that the structural transformation of FL ZnO originates from the competition between the high energy of the O 2pz orbital in the graphitic structure and the polar-surface-induced dipole energy in the wurtzite structure. Our microscopic understanding guides a clear direction of regulating the atomic structure of FL ZnO, further optimizing its electronic properties, which benefits developing function-advanced 2D stacked devices.
Article
Rocksalt-structured (RS) Mg x Zn 1- x O films with x =0.65-1.0 were grown on MgO (100) substrate by the mist chemical vapor deposition method. Comparative study for the RS-Mg 0.92 Zn 0.08 O films grown under slow and rapid-cooling rates apparently showed simultaneous reductions in the surface pit density, full width at half maximum values for the X-ray diffraction peak, and defect-related cathodoluminescence (CL) for the film grown under the slow-cooling rate. CL spectra for the RS-Mg x Zn 1- x O films grown under the slow-cooling rate eventually showed near-band-edge emission peaks in 180-190 nm spectral range for MgO molar fraction x ≥0.92 at room temperature.
Article
Temperature-dependent cathodoluminescence spectra were measured for rock salt-structured MgxZn1−xO films with x = 0.95–0.61. The Mg0.95Zn0.05O film exhibited the shortest deep UV peak wavelength of 199 nm (6.24 eV) at 6 K. Relatively high equivalent internal quantum efficiencies of 0.9%–11% were obtained. The Tauc plots, which were obtained from temperature-dependent optical transmittance measurements, exhibited large Stokes-like shifts of 0.7–0.9 eV at 6–300 K. Time-resolved photoluminescence (PL) signals at 7 K exhibited fast and slow decay components. The fast decay component had PL lifetimes of 2.59–3.08 ns, and the slow decay component far exceeded the measurement time range of 12.5 ns. The fast decay constant reflected the transfer lifetime of the photoexcited carriers to certain trapping centers. These centers were tentatively ascribed to Zn-related isoelectronic trapped-hole centers and may be a cause of the large Stokes-like shifts. The signals at 300 K exhibited very short PL lifetimes of 120–180 ps. The PL lifetimes were mainly attributed to the nonradiative recombination lifetime. Simultaneous decreases in the Zn-related isoelectronic trapped-hole centers and the nonradiative recombination centers were found to be necessary to improve the DUV emission properties of RS-MgxZn1−xO films.
Article
Full-text available
Using self-consistent band-structure methods, we analyze the remarkable anomalies (>50%) in the energy-band gaps of the ternary IB-IIIA-VIA2 chalcopyrite semiconductors (e.g., CuGaS2) relative to their binary zinc-blende analogs IIB-VIA (e.g., ZnS), in terms of a chemical factor ΔEgchem and a structural factor ΔEgS. We show that ΔEgchem is controlled by a p-d hybridization effect ΔEgd and by a cation electronegativity effect DEgCE, whereas the structural contribution to the anomaly is controlled by the existence of bond alternation (RAC≠RBC) in the ternary system, manifested by nonideal anion displacements u-1/4≠0. All contributions are calculated self-consistently from band-structure theory, and are in good agreement with experiment. We further show how the nonideal anion displacement and the cubic lattice constants of all ternary chalcopyrites can be obtained from elemental coordinates (atomic radii) without using ternary-compound experimental data. This establishes a relationship between the electronic anomalies and the atomic sizes in these systems.
Article
Full-text available
We have studied systematically the chemical trends of the band-gap pressure coefficients of all group IV, III-V, and II-VI semiconductors using first-principles band-structure method. We have also calculated the individual “absolute” deformation potentials of the valence-band maximum (VBM) and conduction-band minimum (CBM). We find that (1) the volume deformation potentials of the Γ6c CBM are usually large and always negative, while (2) the volume deformation potentials of the Γ8v VBM state are usually small and negative for compounds containing occupied valence d state but positive for compounds without occupied valence d orbitals. Regarding the chemical trends of the band-gap pressure coefficients, we find that (3) apΓ-Γ decreases as the ionicity increases (e.g., from Ge⃗GaAs⃗ZnSe), (4) apΓ-Γ increases significantly as anion atomic number increases (e.g., from GaN⃗GaP⃗GaAs⃗GaSb), (5) apΓ-Γ decreases slightly as cation atomic number increases (e.g., from AlAs⃗GaAs⃗InAs), (6) the variation of apΓ-L are relatively small and follow similar trends as apΓ-Γ, and (7) the magnitude of apΓ-X are small and usually negative, but are sometimes slightly positive for compounds containing first-row elements. Our calculated chemical trends are explained in terms of the energy levels of the atomic valence orbitals and coupling between these orbital. In light of the above, we suggest that “empirical rule” of the pressure coefficients should be modified.
Article
A change in the banding pattern of the distal end of the third chromosome in Drosophila pseudoobscura has been found. It appears to be produced by homozygosis for a recessive gene, which is called "salivary" (sal) in this report.
Article
This chapter presents an historical overview of high pressure in semiconductor physics. There are two major contributions to the development of apparatus in the post-Bridgman era, which have led to considerable expansion of activity in the field. The first is the introduction of strong, flexible, high-pressure tubing to separate a massive pressure-generating system from the final experimental vessel—which may be at the exit of a spectrometer, between the poles of a magnet, or settled deep in the interior of a cryostat. The second is the advent of the compact diamond anvil cell, which avoids the storage of large energies in compressed fluids. Raman scattering has also been used extensively to map out certain portions of the phase diagram. The pressure dependence of all Raman-active zone-center phonons has been determined as 11 GPa at ambient temperature. Information about zone-edge modes is also found from second- and third-order scattering processes.
Article
A new design for a diamond anvil cell is described. Its main originality is that the force on the piston is generated by pressurized helium, which pushes an annular membrane. It specially permits fine control and adjustment of the force Applied to the anvils, spatial stability of the sample under varying pressure, axial thrust and large optical aperture on both sides of the cell. Selected applications are presented.
Article
We present measurements of ZnO exciton peak energies, E0, at pressures up to 107.3 kbar. Smoke samples consisting of randomly oriented single crystal particles were prepared by oxidizing metallic zinc in air and were collected on one diamond face of a Merrill-Bassett pressure cell. Pressures were measured by the ruby fluorescence technique. In the pressure range between 5 and 90 kbar, our results indicate a consistent linear dependence with dE0/dP = 2.33 × 10−3 eV kbar−1 for both increasing and decreasing pressures. A mixed phase structure is suggested by the observed irregular peak shapes and measured pressure dependence for the sample that had been taken beyond ⋍ 90 kbar where the transformation to the NaCl structure has been reported.
Article
Large single crystals of ZnO (up to 20 g weight) have been grown by oxidation of Zn-vapour. The crystals contain foreign impurities in a concentration of about 1–3 ppm. Furthermore, there have been grown also Li-, Na-, Cu-, Ga-, In- and Mn-doped crystals using this method.
Article
The influence of sample shape, magnetic field direction and magnetic history on the penetration and trapping of magnetic flux in superconducting lead has been investigated ultrasonically.
Article
We report the results of a combined experimental and theoretical investigation on the stability and the volume behavior under hydrostatic pressure of the rocksalt (B1) phase of ZnO. Synchrotron-radiation x-ray powder-diffraction data are obtained from 0 to 30 GPa. Static simulations of the ZnO B1 phase are performed using the ab initio perturbed ion method and the local and nonlocal approximations to the density-functional theory. After the pressure induced transition from the wurtzite phase, we have found that a large fraction of the B1 high-pressure phase is retained when pressure is released. The metastability of this ZnO polymorph is confirmed through the theoretical evaluation of the Hessian eigenvalues of a nine-parameter potential energy surface. This allows us to treat the experimental and theoretical pressure-volume data on an equal basis. In both cases, we have obtained values of the bulk modulus in the range of 160–194 GPa. For its zero-pressure first derivative, the experimental and theoretical data yield a value of 4.4±1.0. Overall, our results show that the ZnO B1 phase is slightly more compressible than previously reported.
Article
We report total energy and electronic structure calculations for ZnO in the B4 (wurtzite), B3 (zinc blende), B1 (rocksalt), and B2 (CsCl) crystal structures over a range of unit cell volumes. We employed both the local-density approximation (LDA) and the PBE96 form of the generalized gradient approximation (GGA) together with optimized Gaussian basis sets to expand the crystal orbitals and periodic electron density. In agreement with earlier ab initio calculations and with experiment, we find that the B4 phase of ZnO is slightly lower in energy than the B3 phase, and that it transforms first to the B1 structure under applied pressure. The equilibrium transition pressure pT1 is 6.6 GPa at the LDA level of theory and 9.3 GPa in the GGA, compared to experimental values around 9 GPa. This confirms a trend seen by other workers in which the LDA underestimates structural transition pressures which are more accurately predicted by the GGA. At much higher compression, we predict that the B1 phase of ZnO will transform to the B2 (cesium chloride) structure at pT2=260GPa (LDA) or 256 GPa (GGA) indicating that gradient corrections are small for this material at megabar pressures. This is the first quantitative prediction of this transition in ZnO, and should be testable with diamond-anvil techniques. We predict that ZnO remains a semiconductor up to pT2. For comparison we find that the B1 to B2 transition in MgO occurs at 515 GPa with either LDA or GGA, in excellent agreement with other ab initio predictions.