You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/343471663

Controlling the thermoelectric power of silicon–germanium alloys in


different crystalline phases by applying high pressure

Article  in  CrystEngComm · August 2020


DOI: 10.1039/D0CE00672F

CITATIONS READS

7 677

4 authors:

Natalia Morozova Igor Korobeinikov


Institute of Metal Physics Institute of Metal Physics
36 PUBLICATIONS   400 CITATIONS    15 PUBLICATIONS   220 CITATIONS   

SEE PROFILE SEE PROFILE

N.V. Abrosimov Sergey V Ovsyannikov


Leibniz-Institut für Kristallzüchtung University of Bayreuth
488 PUBLICATIONS   6,011 CITATIONS    195 PUBLICATIONS   2,762 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Surface preparation and characterization for X-ray crystal optics for metrology and imaging. View project

Advanced Physical Research View project

All content following this page was uploaded by Sergey V Ovsyannikov on 22 September 2020.

The user has requested enhancement of the downloaded file.


Volume 22
Number 33
7 September 2020
Pages 5403–5560

CrystEngComm
rsc.li/crystengcomm

ISSN 1466-8033

PAPER
Natalia V. Morozova et al.
Controlling the thermoelectric power of silicon–germanium
alloys in different crystalline phases by applying high pressure
CrystEngComm
PAPER

Controlling the thermoelectric power of silicon–


Cite this: CrystEngComm, 2020, 22,
germanium alloys in different crystalline phases by
5416 applying high pressure
Natalia V. Morozova, *a Igor V. Korobeinikov, a
Nikolay V. Abrosimovb and Sergey V. Ovsyannikov cd

Silicon, germanium and their alloys are classical semiconductors that play an important role in fundamental
sciences and are the basis for modern microelectronics, optoelectronics, energy conversion, and other
applications. The thermoelectric power (Seebeck coefficient) of p- and n-type materials in combination
with the electrical and thermal conductivities characterizes the efficiency of thermal-to-electric energy
conversion. In this work we experimentally show how one can effectively control the thermoelectric power
of silicon-germanium alloys using an applied high pressure. We measured the Seebeck coefficient and the
electrical resistance for several Si1−xGex crystals with different compositions under applied high pressure in
(i) their original semiconductor cubic-diamond structure, (ii) across a phase transition to a metal phase at
about 9–13 GPa, and (iii) across phase transformations to different metastable phases on pressure release.
These studies were carried out for several pressurization and decompression cycles. The Si–Ge samples
were examined by X-ray diffraction and Raman spectroscopy. After the high-pressure cycling the Si–Ge
samples transformed into two metastable phases, namely, a cubic bc8 phase (Si-III) with a p-type electrical
Received 6th May 2020, conductivity in the Si-dominant samples, and a tetragonal st12 phase (Ge-III) in the Ge-dominant alloys,
Accepted 6th July 2020
whose conductivity type depended on the Si content. The dramatic pressure-driven changes in the
thermoelectric power of Si–Ge crystals we found suggest that these semiconductors are promising for use
DOI: 10.1039/d0ce00672f
in various stress-controlled electronic junctions, such as switches, p–n diode elements, n–p–n (p–n–p)
rsc.li/crystengcomm transistors, and multi-layer heterostructures with alternating types of electrical conductivity.

1. Introduction temperature synthesis routes, a number of novel


polymorphs of Si and Ge can be fabricated, e.g., Si24
Silicon, germanium and their alloys (Si–Ge) are prominent polymorph,22 Si10 polymorph,23,24 hexagonal phases of Ge
semiconductors that play an important role in fundamental and SiGe alloys with direct band gaps,25 and many other
sciences and have various industrial applications. Silicon– phases predicted to be stable.26–33
germanium alloys are promising for their use in fields such Because of low values of the thermal conductivity and
as bipolar technologies,1–3 photonic applications,4,5 and significant values of the Seebeck coefficient at high
lithium-ion batteries,6–9 and in energy conversion temperatures bulk Si–Ge alloys are used in various
technologies, in which, they serve as elements of either solar thermoelectric modules for energy harvesting.11–21 The
cells or thermoelectric devices.10–21 Hence, these materials efficiency of such thermoelectric modules, figure of merit
are always the focus of further investigations. Using (ZT), directly depends on the thermoelectric power (Seebeck
different methods, for example, high-pressure high- effect, S) as ZT = TS 2σ/κ, where T is the temperature, σ is the
electrical conductivity, and κ is the total thermal conductivity,
consisting of lattice and electron contributions. The
a
M. N. Miheev Institute of Metal Physics of Ural Branch of Russian Academy of
thermoelectric power, which shows a magnitude of
Sciences, 18 S. Kovalevskaya Str., Yekaterinburg 620137, Russia. thermoelectric voltage that can be generated in a sample
E-mail: morozova@imp.uran.ru under application of a temperature difference of 1 K, is very
b
Leibniz-Institut für Kristallzüchtung (IKZ), Max-Born-Str. 2, Berlin, 12489, sensitive to changes in crystal and electronic band structures
Germany
c
of this sample. Hence, a possibility to control the value and
Bayerisches Geoinstitut, Universität Bayreuth, Universitätsstrasse 30, D-95447,
Bayreuth, Germany. E-mail: sergey.ovsyannikov@uni-bayreuth.de
sign of the thermoelectric power of materials in general, and
d
Institute for Solid State Chemistry of Ural Branch of Russian Academy of Sciences, of Si–Ge alloys, in particular, would be highly important both
91 Pervomayskaya Str., Yekaterinburg 620219, Russia for thermoelectric technologies themselves and for

5416 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

fabrication of various controls and switches (e.g., p–n switch prone to the formation of Si-rich clusters in the germanium
of conductivity type) for novel nanoelectronic devices. lattice thereby causing both local distortions of the crystal
Like silicon and germanium, their alloys crystallize in a structure and changes in the electron density
cubic-diamond structure (Fig. 1(a) and (b)),34–36 whose lattice configuration.52,53 Meanwhile, the larger Ge atoms are prone
parameter varies from 5.43 Å for silicon to 5.66 Å for to entering the silicon lattice as separate atoms that leads to
germanium with an insignificant deviation from Vegard's the appearance of internal compressive strains.54–59 To
law.34,37–39 Si1−xGex alloys (x > 0.85) have higher mobility compensate these strains the Ge atoms interact with other
values than silicon,40–42 and their electrical conductivity point defects, and often vacancies are formed near the Ge
tends to increase with the Ge content.42 The fundamental atoms.60,61 One can surmise that an applied high pressure/
energy gap of Si1−xGex depends on its composition but in stress could modify the native defect structure of Si–Ge alloys
different manners for Si- and Ge-dominant alloys (Fig. 1(c)). thereby changing their electronic and other properties.
As can be seen in Fig. 1(c) the energy gap in the Si-dominant Since semiconductor elements are the basis of modern
alloys moderately narrows with the Ge content; in contrast, electronics, both novel materials and alternative methods,
in the Ge-dominant alloys the band gap more progressively which could effectively tune their optoelectronic properties,
widens with the Si content from 0.64 eV to about 0.85 eV for are always in much demand.62 Variation in the temperature
the Si0.15Ge0.85 composition (Fig. 1).2,43–47 can induce drastic changes in electronic transport properties
In Si–Ge alloys, which crystallize in the cubic-diamond- (e.g., p–n switching), for example, in Ag10Te4Br3,63 AgBiSe2,64
type structure with the only unique crystallographic site for AgCuS,65 Ag10Te4Br2.6Cl0.4,66 and in LaxSr2−xTiFeO6
67
atoms, the Si and Ge atoms can be randomly distributed over perovskites. However, this approach can only be used to a
a crystal volume.48,49 However, in spite of Si/Ge being limited extent in practical micro- or nanoelectronic devices.
isovalent impurities for each other, because of the sizable Other methods like controlled doping of a surface layer,
difference in their atomic radii (Fig. 1(a) and (b)) these alloys implantation or irradiation and consequent annealing68 are
may be characterized by different point defects depending on not always applicable too. Meanwhile, methods based on
their chemical composition and preparation method. For strain engineering and tuneable applied stress, which are
minor Si/Ge substitutions one can expect a uniform chemical capable of changing the properties in micro- or nanoscale
composition over a sample, whereas the higher levels of Si/ regions of semiconductors,69 seem more applicable to
Ge substitutions can lead to partial Si/Ge segregation and the different optoelectronic devices, e.g., in switches, diodes,
formation of regions that are richer either in Si or in Ge.50,51 transistors, or memory elements. A number of academic
In some work it was noted that the smaller Si atoms are more studies of semiconductors reported the observation of
remarkable pressure-induced structural or electronic
transitions that are concurrent with drastic changes in their
optoelectronic properties, including p–n inversions of the
dominant conductivity type.70–75 These findings corroborated
the high potential of this method.
In this work we investigated the effect of an applied
high pressure on the thermoelectric power of silicon–
germanium alloys with Si- and Ge-dominant compositions,
Si1−xGex (x = 1, 0.98, 0.93, 0.87, 0.026, and 0), in the
pressure range up to 17 GPa, that is, across a
semiconductor–metal phase transition to a metal β-Sn-type
phase at about 9–13 GPa.76–100 We found that the applied
pressure can dramatically change the value of the Seebeck
coefficient in the cubic-diamond structure of Si–Ge alloys,
and for some compositions it can even invert the sign of
the thermopower. High-pressure cycling across the
semiconductor–metal phase transition point76–100 allowed
stabilization of the metastable phases of Si1−xGex with
different crystal structures and different electronic
properties. The crystal structures of Si1−xGex alloys were
examined by X-ray diffraction and Raman spectroscopy. The
Fig. 1 The cubic-diamond crystal structure of (a) germanium slightly findings of our work suggest that an applied pressure can
substituted with silicon, SixGe1−x and (b) silicon slightly substituted with directly affect the competition of different conductivity
germanium, Si1−xGex. (c) Composition dependence of the fundamental
mechanisms in Si–Ge alloys. These results are of interest
indirect energy gap of Si–Ge alloys replotted from the literature data
for relaxed solid solutions (open circles connected by a solid curve,
from both fundamental and applied perspectives. Based on
from ref. 43). The red arrows correspond to the compositions we the results of our work we proposed several potential
investigated in this work. innovative applications of Si–Ge alloys.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5417
Paper CrystEngComm

2. Experimental details semispherical cavities in the centre. This cell allowed


2.1. Synthesis of the Si–Ge samples measurements of bulk ‘3D’ samples with typical sizes of
about 150 × 150 × 150 μm3 under high pressure up to 9 GPa
Bulk crystals of Si0.974Ge0.026 were grown by the Czochralski (Fig. 2(a)).112 Another cell was of a modified Bridgman type,
technique and characterized by conventional methods.101 and was equipped with sintered-diamond anvils having flat
These crystals had a p-type electrical conductivity and a low working surfaces of 600 μm in diameter.113,114 In this cell we
hole concentration of about 5 × 1015 cm−3.102 Bulk crystals of could measure thin ‘2D’ samples with typical sizes of about
pure germanium with n- and p-type conductivities and low 200 × 200 × 30 μm3 in the higher pressure range up to 20
carrier concentrations of about 1014 cm−3 were provided by GPa.112 The high-pressure cell with semispherical cavities in
different suppliers. Large-grain polycrystalline Ge-dominant the anvils provided a quasi-hydrostatic compression regime
alloys, Si0.02Ge0.98, Si0.07Ge0.93, and Si0.13Ge0.87 were for a sample.115,116 In another cell, which was equipped with
synthesized under high-pressure high-temperature conditions conventional flat anvils, the ratio of the sample thickness (30
in a multi-anvil press.103 The synthesis procedures were the μm) to the culet size of the anvils (600 μm) amounted to
same as reported in previous studies.104–106 For the high- about 0.05. This value is less than the threshold value of
pressure electronic transport measurements we selected 0.055–0.057, which was found earlier to correspond to the
microscopic single-crystal grains of these Ge-dominant Si–Ge suppression of uniaxial pressure components in solid
alloys with diameters exceeding ∼150 μm. The compositions media.117 In both cells a sample was loaded into a hole
of the samples were controlled by scanning electron drilled in the centre of a toroidal-shaped container made of
microscopy (SEM) using a LEO-1530 instrument. For the limestone mineral (predominantly CaCO3).114,118 This sample
exception of Si0.13Ge0.87, which exhibited minor fluctuations container also served as a gasket and as a pressure-
in the composition over the volume, the other samples were transmitting medium. Limestone has a low yield strength
chemically uniform. shear, and hence, it is a suitable material for pressure-
transmitting medium,119 which may be comparable with
2.2. Characterization of the Si–Ge samples by Raman and NaCl powder.120 The main component of the limestone
X-ray diffraction studies (calcite) is known to undergo pressure-driven phase
Both the as-grown and recovered samples after the high-
pressure experiments were examined by Raman spectroscopy
and X-ray diffraction. For the Raman studies we utilized a
LabRam spectrometer and a He–Ne laser (λ = 632.8
nm).107,108 A spectral resolution of the setup amounted to
about 0.5 cm−1. The structural examinations were carried out
using high-brilliance Rigaku diffractometers with λ = 0.5594
Å (Ag Kα X-ray source) and 0.7108 Å (Mo Kα X-ray source). For
one sample, namely, Si0.974Ge0.026, we performed in situ
Raman spectroscopic experiments under applied high
pressure to 40 GPa. These Raman studies were accomplished
both for pressurization and decompression runs and allowed
us to document the formation of metastable polymorphs
upon pressure release. For this work we employed a diamond
anvil cell of a piston-cylinder-type (BX90)109 equipped with a
rhenium gasket. A hole in the rhenium gasket, in which the
samples were placed, was filled with a neon pressure-
transmitting medium using a BGI's gas loading apparatus.110
The neon medium provided a high-quality quasi-
hydrostaticity around the sample up to the maximum
pressure of 40 GPa.111 The pressure values were determined
by the shift of the ruby luminescence line. Fig. 2 (a) A simplified scheme of the high-pressure cell with
semispherical cavities in the anvils that was used for the thermoelectric
measurements. 1 – Sample, 2 – sample container made of limestone, 3
2.3. Characterization of the Si–Ge samples by electronic
– anvils, 4 – supporting plungers. The electrical probes show how the
transport studies thermoelectric voltage (ΔU) and the temperature difference (ΔT) were
Measurements of the thermoelectric power (Seebeck measured. (b) Examples of the Seebeck coefficient determination from
linear slopes, as S = −ΔU/ΔT at fixed pressures for different Si–Ge
coefficient) and electrical resistance under applied high
alloys. 1 – The metastable Ge-III-type phase of Si0.02Ge0.98 at about 0
pressure were carried out in two different anvil-type high- GPa, 2 – Si0.974Ge0.026 at about 0 GPa after recovery from 2 GPa, 3 –
pressure cells. One cell was of a toroidal type; its anvils had a Si0.974Ge0.026 at about 10 GPa, 4 – Si0.13Ge0.87 at about 0 GPa, 5 –
working diameter of ∼1 mm and were characterized by Si0.07Ge0.93 at about 0 GPa.

5418 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

transitions at about 1.5–2 GPa, concurrent with notable For the maximal Si content of 13 at% the difference in the
volumetric effects.121 This circumstance also contributed to wavenumbers with pure Ge amounted to only about 2 cm−1
the improvement of quasi-hydrostaticity of the pressure (Fig. 3(a)). This observation was in line with previous Raman
around a sample. Thus, both high-pressure cells equipped studies on Si1−xGex.40,127–138 The other very weak peaks we
with the limestone sample container could provide a quasi- documented in the spectra of the Ge-dominant samples at
hydrostatic compression regime. 354, 385, and 403 cm−1 (Fig. 3(b)) are not observed in the
The high-pressure cells were loaded into an automated spectra of pure germanium, and hence, they could originate
mini-press setup, which smoothly generated a force applied from some Si–Ge-related vibrations. Previous Raman
to a cell with assistance of electric motors, and automatically examinations of Si–Ge alloys with intermediate compositions
recorded all relevant parameters.122 The measurements were also found a peak linked to Si–Ge vibrations about 400 cm−1
performed both for pressurization and decompression runs and noted its sensitivity to many factors, such as
for multiple pressure cycles. The values of the force applied composition, sample morphology and crystallinity.40,134–142
to a cell were automatically measured using a digital Compared to pure silicon, in the Si0.974Ge0.026 composition a
dynamometer, and then, they were converted into pressure wavenumber of the strongest LO phonon softens a bit to 519
values.112 Typical times of pressurization or decompression cm−1. Very weak and broad peaks in the spectra of
cycles amounted to 15–20 minutes. Electrical resistance Si0.974Ge0.026 at 375 and 399 cm−1 (inset in Fig. 3(b)) have a
values were measured by a quasi-four-probe method (two correspondence with the above weak peaks in the Ge-
bifurcated probes).72,123,124 For the thermoelectric power dominant samples. The peaks at 301 and 430 cm−1 we
measurements an upper anvil was electrically heated to observed in Si0.974Ge0.026 were also reported for pure silicon,
generate a temperature difference (ΔT) of several Kelvins and they were assigned in the literature to the transverse
between the upper and lower sides of the sample (Fig. 2(a) acoustical phonon modes, 2TA(X) and 2TA(Σ),
).125 This ΔT difference generated a thermoelectric voltage respectively.79,137,139,143–147
(Fig. 2(b)), and it was measured using thermocouples We investigated the high-pressure behaviour of the
attached near the anvil tips (Fig. 2(a)).126 Seebeck coefficient of the Si–Ge alloys in two different high-
pressure cells. One cell could generate high pressures up to 9
3. Results GPa that corresponded to a stability region of the original
cubic-diamond phase (Si-I, Ge-I) (Fig. 4). The other cell could
The Raman spectra of the Si–Ge alloys are depicted in Fig. 3. generate high pressures somewhat above the semiconductor–
The main LO phonon mode of germanium in the Ge- metal phase transition point about 9–13 GPa,76–96 and
dominant samples only slightly softens with the Si content. allowed measurements of the Seebeck coefficient of both
metal phases with a β-Sn-type structure (Si-II, Ge-II) and the
metastable phases that appear upon pressure release
(Fig. 4).78,84,86,95,147–164 Below, we report the results of our
studies of the high-pressure behaviour of the thermoelectric
power and electrical resistance of the Ge-dominant Si–Ge
alloys (Fig. 5–8) and of Si0.974Ge0.026 (Fig. 9 and 10). We
supplement these results by high-pressure Raman
spectroscopic data for Si0.974Ge0.026 (Fig. 11 and 12) that
documented the phase transitions in pressurization and
decompression cycles.

3.1. Ge-Dominant Si–Ge alloys: high-pressure effect on the


thermoelectric power in the cubic-diamond phase
The pressure dependencies of the thermoelectric power of
the Ge-dominant Si–Ge samples for several pressure cycles
are shown in Fig. 5. Under normal conditions only pure Ge
was characterized by a negative Seebeck coefficient of −330
μV K−1 (Fig. 5(a)). This fact suggested a dominant n-type
electrical conductivity. The germanium substitution with Si
(Si0.02Ge0.98, Si0.07Ge0.93, and Si0.13Ge0.87 samples) led to the
stabilization of the p-type electrical conductivity under
normal conditions (Fig. 5(b and c)). This charge-carrier
Fig. 3 Raman spectra of the Si–Ge alloys under ambient conditions.
The insets in (a and b) show the selected magnified parts of these
inversion could result from the appearance of intrinsic point
spectra. A photograph of the bulk Si0.974Ge0.026 single crystalline defects in the cubic-diamond crystal structure of Ge.165,166 An
sample is given in (b). increase in the Si content led to a decrease in the Seebeck

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5419
Paper CrystEngComm

Fig. 4 Selected crystal structures of silicon and germanium. (a) Under ambient conditions both silicon and germanium crystallize in the cubic-
diamond-type structure with the Fd3̄m space group, #227 (known as Si-I and Ge-I). (b) Under applied high pressure above 9–13 GPa this diamond-
type structure transforms into a metal β-Sn-type structure with the I41/amd space group, #141 (known as Si-II and Ge-II). (c) On decompression
the β-Sn-type structure of Ge usually transforms into a metastable st12 structure with the P43212 space group, #96 (known as Ge-III), which can
be quenched under ambient conditions. (d) On decompression the β-Sn-type structure of Si firstly transforms into a metastable r8 structure with
the R3̄ space group, #148 (known as Si-XII), and (e) then, on pressure release, transforms into a metastable bc8 structure with the Ia3̄ space group,
#206 (known as Si-III). Rarely, depending on stress conditions, after pressure release Ge can also adopt the same structure as Si-III (e) (known as
Ge-IV).

Fig. 5 Pressure dependencies of the thermoelectric power of the Ge-dominant Si–Ge alloys at 295 K under high-pressure cycling below the
semiconductor–metal phase transition point (about 10 GPa). (a) n-Type Ge. (b) Si0.02Ge0.98. (c) Si0.07Ge0.93. (d) Si0.13Ge0.87. The thin arrows indicate
the directions of pressure variation.

coefficient, from about +465 μV K−1 for x = 0.02 to about +200 ‘light’ holes bands followed by a carrier transfer to the ‘light’
μV K−1 for x = 0.13 (Fig. 5). At the initial compression of 1 one, in which the holes are much more mobile.73 Thus, the
GPa the sample of pure Ge irreversibly inverted the electrical present results confirm these previous findings.
conductivity type, and on the successive pressure cycles it The entire pressure dependencies of the thermopower
behaved already as a p-type semiconductor (Fig. 5(a)). An n–p plotted in Fig. 5 corresponded to the stability range of the
inversion of the electrical conductivity type in pure Ge original cubic-diamond-type phase. All the samples
samples was already documented in our previous work.73 demonstrated the apparent kinks in the curves about 2–5
This effect was attributed to a splitting of the ‘heavy’ and GPa for both pressurization and decompression cycles

5420 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

thermopower curves about 2–5 GPa shifted a bit to higher


pressures with the cycle number likely because of a minor
deterioration of the samples (Fig. 5). After these high-
pressure-cycling experiments all the recovered samples were
characterized by nearly identical magnitudes of their Seebeck
coefficients of about +(170–200) μV K−1 (Fig. 5).

3.2. Ge-Dominant Si–Ge alloys: variation in the


thermoelectric power across the phase transitions under
pressure
The pressure dependencies of the thermoelectric power of
the p-type Ge-rich Si–Ge alloys measured across the phase
transition to the metal β-Sn structure (Ge-II) are shown in
Fig. 7. In the pressure range from 1 to 8 GPa these
thermopower curves for the first pressurization cycle were
quite similar to those measured in the toroidal high-pressure
cell are and shown in (Fig. 5(b–d)). On the first pressurization
cycle all the samples conserved the original p-type electrical
conductivity (Fig. 7). In the vicinity of the semiconductor–
metal phase transition about 9–13 GPa,76,77,83–92 the
Fig. 6 Comparison of the pressure behaviour of the thermoelectric
thermopower curves exhibited a distinct turning point (insets
power of the Ge-dominant Si–Ge alloys at 295 K from Fig. 5(a) for the in Fig. 7(c) and (d)). Above the phase transition point in the
first pressurization cycle and (b) for the second pressurization and β-Sn-type structure the Seebeck coefficients of these samples
decompression cycles. (a) Above 1 GPa all the samples demonstrate slightly increased with pressure.
rather similar dependencies of the Seebeck coefficients under
Interestingly, on the first decompression cycle from the
pressure. (b) All the curves exhibit a drop about 3–6 GPa marked by
the bulk arrow for Si0.02Ge0.98. The thin arrows indicate the directions
metal β-Sn phase the thermopower curves for these samples
of pressure variation. demonstrated very different dependencies. Thus, for pure
and slightly-substituted samples (Ge and Si0.02Ge0.98) we
observed an abrupt change in the Seebeck coefficients at 2–3
(Fig. 5). These features likely result from some reversible GPa to high negative values of about −(250–300) μV K−1
reconstruction in their electronic band structures. (Fig. 7(a) and (b)). This behaviour suggested the formation of
Remarkably, the samples with different Seebeck coefficients, a phase with dominant n-type conductivity. The other two
ranging from −330 μV K−1 to +465 μV K−1 at normal pressure, samples, namely, Si0.07Ge0.93 and Si0.13Ge0.87, showed an
demonstrated rather a similar pressure behaviour of the opposite tendency, and below 2 GPa their Seebeck
thermopower above 1 GPa. For example, the thermopower coefficients jumped to high positive values of about +(150 ÷
curves for n-type Ge and p-type Si0.07Ge0.93 almost coincided 170) μV K−1 (Fig. 7(c) and (d)). On the re-pressurization cycles,
above 1 GPa (Fig. 6(a)). These findings strongly suggested Ge and Si0.02Ge0.98 demonstrated a similar behaviour. Their
that all the samples were characterized by an intrinsic thermopower curves showed very smooth bends about 5 GPa
semiconductor conductivity, and the pressure behaviour of and an apparent phase transition to the metal β-Sn phase at
their Seebeck coefficients was tightly linked to the Ge-like about 9 GPa, i.e., at nearly the same pressure as the
band structures of these alloys.2,43–46 Thus, we can conclude transition detected on the first cycle. Furthermore, this phase
that the slightly compressed samples of pure and Si- transition was accompanied by a reversible n–p inversion
substituted germanium are always p-type conductors (Fig. 7(a) and (b)). In contrast, the Seebeck coefficient of the
irrespective of the conductivity type and crystallinity of the Si0.07Ge0.93 sample showed a dependence on the cycle
original samples. A moderate compression above 5–6 GPa number and demonstrated the alternating type of the
strongly suppressed the thermoelectric power of these Ge- dominant electrical conductivity (Fig. 7(c)). In sample with
dominant alloys (Fig. 6). Since the value of the indirect band the highest Si content (Si0.13Ge0.87) the p-type conductivity
gap of germanium was found to widen insignificantly with was conserved in the whole pressure range for all pressure
pressure,167–169 the suppression of the Seebeck coefficient cycles. Note here that the above discussed pressure cycling
was likely caused by an enhancement of n-type contribution up to 8 GPa (Fig. 5), i.e., below the phase transition
that led to a nearly compensated electronic state (i.e., with point,76,77,83–92 did not result in such dramatic variations in
nearly equal partial conductivities of electrons and holes) the Seebeck coefficients. Therefore, these changes should be
(Fig. 6). Starting from the second pressure cycle, the linked to the formation of the metastable phases on the
pressure-driven changes in the thermopower curves of these decompression pathway from the metal β-Sn
samples were reversible (Fig. 5). The bends in the phase.84,85,92,150–156,162,163 Two metastable phases of

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5421
Paper CrystEngComm

Fig. 7 Pressure dependencies of the thermoelectric power of the Ge-dominant Si–Ge alloys at 295 K across the phase transition to the metal
β-Sn-type structure (Ge-II) about 10 GPa for selected pressure cycles for (a) p-type Ge, (b) Si0.02Ge0.98, (c) Si0.07Ge0.93, and (d) Si0.13Ge0.87. The thin
arrows indicate the directions of pressure variation. After the first decompression cycle, the metastable Ge-III-type phase was formed. Hence, the
second and following pressure cycles should correspond to the Ge-III → Ge-II phase transition. The bulk arrows indicate anomalies, which should
be linked to the phase transitions. The insets in (c) and (d) show the magnified parts of the thermopower curves, which demonstrate these
anomalies more distinctly.

germanium, which can be formed after different high- pure Ge we found the unit cell parameters and the atomic
pressure treatments, were reported in the literature. One of coordinates to be as follows: a = 5.933(7) Å, c = 6.966(0) Å, V =
them has a tetragonal st12 structure (Ge-III) (Fig. 4(c)) and is 245.26(7) Å3, Z = 12, Ge1(0.0934, 0.0934, 0), and Ge2(0.1696,
often formed after slow decompression runs.84,85,163,164 The 0.3439, 0.2579) (Fig. 4(c)). These structural data were similar
other metastable phase adopts a cubic bc8 structure (Ge-IV) to those reported previously in the literature.158
and may be quenched to ambient conditions during very fast The significant values of the Seebeck coefficients of about
decompression cycles (less than 1 second).154 −(250–300) μV K−1 in the metastable tetragonal st12 phase
To clarify the origin of the thermopower variations (Ge-III) of Ge and Ge-rich samples (Fig. 7) pointed out their
observed on the decompression cycles (Fig. 7) we examined semiconducting character. A band gap value in this Ge-III
the samples recovered under ambient conditions after these phase may be comparable with a gap of 0.67 eV in the cubic-
experiments (Fig. 8(a)) by Raman spectroscopy and X-ray diamond structure of Ge.73 This finding agreed well with a
diffraction (Fig. 8(c) and (b), (d)). Because of a notch filter recent experimental determination of an indirect band gap of
the Raman spectra below 200 cm−1 were cut, and we could 0.59 eV in the pure Ge-III phase164 as well as with an earlier
detect only three peaks at 226, 242 and 270 cm−1 in all the predicted band-gap value of ∼0.7 eV.152 Our thermopower
samples (Fig. 8(c)). We noted that with the Si content the data indicated that a native defect structure in the Ge-III
wavenumbers of these phonons slightly shifted to higher phase of pure and slightly substituted samples corresponds
frequencies. In the literature these phonon modes were to the n-type electrical conductivity (Fig. 7(a) and (b)). The
assigned to the metastable tetragonal st12 structure (Ge- higher silicon content strongly shifted the conductivity in the
III).73,155,156,170–176 A certain broadening in these spectra may Ge-III phase to p-type (Fig. 7(c) and (d)). For Si0.13Ge0.87 this
be attributed to a minor disorder in the crystal structure led to the stabilization of the p-type conductivity in all the
owing to the pressure cycling over the phase transition phases. For the intermediate composition, Si0.07Ge0.93, the
region.177 In the X-ray diffraction studies this conjecture has thermopower fluctuations on the decompression cycles
been confirmed for all the Ge dominant samples pointed out a strong competition between the n- and p-type
(Fig. 8(b) and (d)). Although, a quality of these structural data partial conductivities (Fig. 7(c)). Under high pressure this
was rather limited we could perform a full-profile analysis of metastable Ge-III phase should also transform into the
these patterns and confirmed that they all can be fit well by thermodynamically stable metal β-Sn-type phase, but the
the P43212 tetragonal space group (#96) of the Ge-III phase. transition pressure remains a point of issue. On the one
An example of these fits for pure Ge is given in Fig. 8(d). For hand, the transition pressure can correspond to the region of

5422 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

Fig. 8 The metastable Ge-III-type phase in the Ge-dominant Si–Ge alloys and its characterization at 295 K. (a) Photograph in transmitted light of a
limestone sample container (it is labelled as 2 in Fig. 2(a)) with one of the Si–Ge samples recovered from the high-pressure experiments. (b)
Example of an azimuth X-ray diffraction image in the full range of 0–360° for the Si0.07Ge0.93 sample. This corresponds to the 1D diffraction pattern
of this sample shown below in (d). (c) Raman spectra of the samples recovered from the high-pressure-cycling experiments to about 17 GPa
(Fig. 7). The dashes indicate the wavenumbers reported for the Ge-III phase in ref. 156. (d) X-ray diffraction patterns for the same samples as in (c).
The Ge-III-type crystal structure of the samples was confirmed by Rietveld refinement. For pure Ge a simulated pattern (thick solid line) together
with the calculated reflection positions (dashes) for this structure is given. The asterisks in the pattern for the Si0.13Ge0.87 sample point out peaks
related to the remains of the limestone container. (e) Pressure dependencies of the electrical resistance for the Si–Ge samples across the Ge-III →
Ge-II phase transition (marked by the bulk arrow on the sample of Ge). (f) Pressure dependencies of the thermoelectric power of the Si–Ge
samples across the Ge-III → Ge-II phase transition, from Fig. 7. The bulk arrows indicate the direct and return phase transitions. The thin arrows in
(e) and (f) indicate the directions of pressure variation.

the thermodynamic stability of the β-Sn phase, i.e., it may 3.3. Si-Dominant Si–Ge alloys: high-pressure effect on the
be the same as in the cubic-diamond-type Ge.76,77,83–92 But thermoelectric power in the cubic-diamond phase
on the other hand, since this metastable phase is already
strongly ‘pre-compressed’ by about 10–12%, compared to The silicon-rich Si1−xGex crystal (Si0.974Ge0.026) had a p-type
the cubic-diamond one,84,85,148,174 less applied pressure is electrical conductivity and showed a high value of the Seebeck
needed to increase the density of the materials to a coefficient above 1 mV K−1 (Fig. 9(a)). We carried out the
critical level to cause this transition. Our results measurements using high-pressure cells equipped with
demonstrated that the Ge-III → Ge-II phase transition pressed electrical probes, and for this reason we could not
occurs rather abruptly and its transition pressure tends to measure the thermopower without applied pressure, i.e.,
decrease with the Si content (Fig. 8(e) and (f)). These under ambient conditions. We observed that application of
findings hint that even the metal phase of the Si–Ge even a minor pressure of less than 1 GPa can spectacularly
alloys could be stabilized in thin strained films. Thus, our change the electronic transport properties of this crystal
work showed that the thermoelectric power of the Si–Ge (Fig. 9). An electrical conductivity of Si0.974Ge0.026 was strongly
alloys may be dramatically changed by using an applied enhanced to 1 GPa by about three orders of magnitude (the
high pressure. These thermopower changes are linked not inset in Fig. 9(a)), and the dominant conductivity type was
only to the structural phase transitions but to different changed from p- to n- (Fig. 9). Pure silicon did not exhibit
electronic states, which may be varied by using an applied such effects on the electrical resistance178 and the
pressure. thermopower.179 We found that the pressure-driven drop in

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5423
Paper CrystEngComm

Fig. 9 Pressure dependencies of the thermoelectric power for two samples (#1 and #2) of Si0.974Ge0.026 below the semiconductor–metal phase
transition point at 295 K. The thin arrows indicate the directions of pressure variation. The inset in (a) shows a pressure dependence of the
electrical resistance for sample #1 for two pressurization cycles. The insets in (b) show the pressure and time dependencies of the thermopower
for sample #2 on relaxation release from a pre-compression run to 0.75 GPa. On this decompression a return n–p sign inversion of the Seebeck
coefficient occurs about 0.55 GPa.

Fig. 10 The formation of the metastable Si-III phase in Si0.974Ge0.026 and its characterization at 295 K. (a) Pressure dependencies of the
thermoelectric power of sample #3 of Si0.974Ge0.026 across the phase transition to the metal β-Sn-type phase (Si-II) about 11–13 GPa for three
pressure cycles. The thin arrows indicate the directions of pressure variation. After the first decompression run Si0.974Ge0.026 transformed into the
metastable Si-III-type phase. The inset in (a) depicts a pressure dependence of the electrical resistance of this sample across the Si-III → Si-II phase
transition. (b) Magnified pressure dependencies of the thermoelectric power from (a), which demonstrate smooth bends in the region of the phase
transition to the metal β-Sn phase about 11 GPa. The decompression cycles exhibit a dip about 4 GPa and a minimum at ∼0.6 GPa (better shown
in the inset), which could be related to the phase transitions to the metastable phases Si-XII and Si-III, respectively. The second and third
pressurization cycles show a feature about 2 GPa that could be linked to the Si-XII → Si-III phase transition. (c) Raman spectrum of the metastable
Si-III phase of Si0.974Ge0.026 collected from the sample recovered from the high pressure experiments to 17 GPa (a and b). The wavenumbers are
given near the peaks. Data from the literature for Si-III phase, based on the studies by OJ156 and HS,155 are shown by dashes. (d) X-ray diffraction
pattern of the metastable Si-III phase of Si0.974Ge0.026 collected from the sample recovered from the high-pressure cycling experiments to about
17 GPa (a and b). The Si-III-type crystal structure was confirmed by Rietveld refinement; a simulated pattern is shown by a solid line, and the
calculated reflection positions are shown by dashes. The inset in (d) shows a quarter of a 2D X-ray diffraction image of this sample.

the electrical resistance value is reversible and reproducible Meanwhile, the p–n inversion at 0.5–1 GPa may be reversible
for the successive pressure cycles (inset in Fig. 9(a)). only if the maximal applied pressure does not exceed

5424 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

Fig. 11 (a) Photographs of a central part of a high-pressure diamond anvil cell with the Si0.974Ge0.026 sample (marked by a bulk blue arrow) at
selected pressures on pressurization run. A ruby ball in the Centre was used for pressure determination. Above 16 GPa Si0.974Ge0.026 demonstrates
a metallic lustre consistent with the Si-I → Si-II phase transition.76–80,86–88,93–96,155–157 Raman spectra evolution for the Si0.974Ge0.026 crystal on (b)
pressurization and (c) decompression cycles indicated the semiconductor–metal phase transition between 13 and 16 GPa, and the transitions to
the metastable structures on pressure release. The asterisks mark the parasite peaks. The pressure values are given near the selected spectra. (b)
The inset shows the magnified parts of the Raman spectra in the metal phase. The spectra collected at 16 and 16.9 GPa correspond to the Si-II
phase, whereas the one at 18 GPa and above – to the Si-V phase, which has no Raman active modes.87 (c) The metastable phase Si-III was
quenched under ambient conditions. The wavenumbers in the spectrum of Si-III are given near the peaks. They were consistent with previous
studies.155,156

approximately 1 GPa. For instance, we compressed one 3.4. Si-Dominant Si–Ge alloys: variations in the
Si0.974Ge0.026 crystal to 0.75 GPa, and then slightly released the thermoelectric power across the phase transitions under
force of the press. This led to relaxation changes in both the pressure
pressure value in the high-pressure cell and the electronic
properties of the sample. Thus, for 40 minutes the pressure in One sample of Si0.974Ge0.026 (#3) cut from the same bulk
the cell gradually decreased to 0.45 GPa. This caused a return single-crystalline ingot was subjected to pressure cycling up
n–p inversion about 0.55 GPa (insets in Fig. 9(b)). However, if to 17 GPa across the semiconductor–metal phase transition
the applied pressure well exceeds 1 GPa (this approximately (Fig. 10(a)). This sample #3 also exhibited a p–n inversion
corresponds to the negative extremum on the curves) the after the start of the compression, but its thermopower
original p-type conductivity can no longer be easily restored variations looked more smeared, compared to those observed
upon decompression. The p–n inversion in Si0.974Ge0.026 is likely in samples #1 and #2 (Fig. 9). This difference can be
linked to a modification of its native defect structure, and explained by the fact that the measurements of samples #1
hence, the return n–p inversion could be probably stimulated and #2 were conducted in the toroidal high-pressure cell that
by sample annealing. In one sample of Si0.974Ge0.026 cut from provided better quasi-hydrostaticity, compared to the higher-
the same crystal we observed even more spectacular effects pressure cell with flat anvils, which was utilized for the
consisting of a triple p–n–p–n sign inversion upon compression measurements of sample #3. Similar to samples #1 and #2,
to 1.5 GPa (Fig. 9(a)). These anomalies were likely linked to sample #3 also changed into a strongly compensated
minor pressure gradients in the high-pressure cell that could electronic state above a few GPa (Fig. 10(a)). A gentle bend on
spontaneously appear in the solid pressure-transmitting the thermopower curves for all pressurization cycles in the
medium across the sample. A stronger compression of vicinity of 11 GPa (Fig. 10(b)) can be attributed to the phase
Si0.974Ge0.026 crystals resulted in an electronic transition to a transition to the metallic β-Sn phase (Si-
compensated state that was characterized by nearly equivalent II).76–80,86–88,93–96,155–157 On the decompression cycles we
partial conductivities of holes and electrons and close to zero found a steeping dip on the curves about 4 GPa and a reverse
thermopower values (Fig. 9). This compensated state was rather jump at 0.6 GPa (Fig. 10(b)). These features could be linked
robust and preserved on pressure release. In this case, a to the formation of the metastable phases, namely, the
resulting value of the Seebeck coefficient was very small and rhombohedral phase, r8 (Si-XII) and the body-centred cubic
could be either positive or negative (Fig. 9). Generally, the effects phase, bc8 (Si-III) (Fig. 4), although, the higher transition
we found suggest that Si1−xGex crystals with a low Ge content pressures for these phases were previously reported.88,158,159
are highly promising for different piezoelectric and Note here that the difference in the decompression pathways
optoelectronic technologies. for Si and Ge was explained in the literature by the fact that

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5425
Paper CrystEngComm

except the low-frequency region. In particular, we did not


observe a reported earlier peak at 182 cm−1 likely because of
a significant suppression of this region using a notch filter.
At the same time, we observed a new peak at 294 cm−1
(Fig. 10(c)), which was absent in the spectra of pure
Si.155–157,160,161,180–182 Note here that in the in-situ high-
pressure Raman spectroscopic studies of this sample that are
discussed below a similar peak was also observed (Fig. 11(c)).
The X-ray diffraction studies of this sample confirmed the
Ia3̄ space group (#206) and found the unit-cell parameters as
a = 6.623(4) Å, V = 290.56(3) Å3, and Z = 16 (Fig. 4(e)). These
results agreed well with the data of a recent high-quality
study on Si-III polymorphs.183
Thus, we can conclude that both the rhombohedral phase,
r8 (Si-XII), and the body-centred cubic phase, bc8 (Si-III), are
p-type conductors with a nearly semimetal conductivity
(Fig. 10(b)). A phase transition from these metastable phases
to the metallic β-Sn structure (Si-II) was not accompanied by
any dramatic changes in the transport properties, as seen
from the thermopower and the electrical resistance data
(inset in Fig. 10(a)). These findings are in line with the
thermopower data for pure Si.179 A previous study also
reported for the Si-III phase a p-type semimetal conductivity
with a hole concentration of 5 × 1020 cm−3.149 Following
investigations reported a semiconducting-like dependence of
the electrical resistivity at normal pressure for the Si-III
phase.183 More recent work discovered an ultra-narrow direct
band gap of 30 meV in this phase.184

3.5. Si-Dominant Si–Ge alloys: Raman spectroscopy under


high pressure
To examine the crystal structure stability of the Si–Ge
Fig. 12 Pressure dependencies of the wavenumbers of the samples we carried out a Raman spectroscopic study on the
Si0.974Ge0.026 crystal for pressurization and decompression cycles. (a)
sample of Si0.974Ge0.026 (Fig. 11 and 12). We found that the
The wavenumber of the main LO phonon exhibits a kink in its pressure
curve at 2 GPa. The inset shows the intensity variation of this LO semiconductor–metal phase transition in Si0.974Ge0.026 occurs
phonon peak under pressure. (b) The phonon at 430 cm−1, which is between 13 and 16 GPa (Fig. 11(b)), i.e., at somewhat higher
shown in the inset and assigned to the 2TA(Σ) silicon mode, also pressures than in pure bulk Si.76–80,86–88,93–96,155–157 This
exhibits a kink at 2 GPa. It seems that this phonon peak probably splits finding was in line with previous Raman studies of Ge-
above 3 GPa. (c) Wavenumbers of the other phonon modes of Si-I, Si-
substituted silicon.96 Tentatively, this increased transition
II, Si-III, and Si-XII phases. The inset illustrates the phase transition
between the metastable phases, Si-XII → Si-III on pressure release. pressure can be explained by a minor lattice expansion due
to the incorporation of bigger Ge atoms (Fig. 1(b)).7,34,185–187
In visual examinations through a diamond window we
observed the appearance of a distinct metallic lustre in the
the kinetic barriers for the different phase transitions in Si sample at a pressure of 16 GPa (Fig. 11(a)). In the metal
and Ge are not the same, and upon decompression from the phase of Si0.974Ge0.026 we found a doublet at 385 and 405
β-Sn phase each of these elements follows the kinetically cm−1 in a spectrum collected at 16 GPa and a single peak at
simplest pathway instead of a return transformation to the 413 cm−1 in a spectrum collected at 16.9 GPa (inset in
more energetically-favourable cubic-diamond structure.162 Fig. 11(b)). At higher pressures up to 39 GPa no clear Raman
To verify that sample #3 of Si0.974Ge0.026 recovered after peaks were detected. These results generally agreed with
the pressure-cycling experiments up to 17 GPa previous studies for Si.87 This work also reported the
(Fig. 10(a) and (b)) indeed transformed into the metastable appearance of a few phonon modes with similar
Si-III-type phase we examined it by Raman spectroscopy and wavenumbers in the β-Sn-type phase of silicon (Si-II) and the
X-ray diffraction (Fig. 10(c) and (d)). A Raman spectrum of loss of the Raman spectra above 16 GPa.87
this sample looked nearly identical to those reported in the On a decompression cycle we observed the formation of
literature for pure Si in the Si-III phase155–157,160,161,180–182 the metastable high-pressure phases (Fig. 11(c)). The Raman

5426 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

     
spectra of these phases were similar to those reported k σ n − σ p Eg 5 σn 5 σp 3 m*p
S¼ − × þ rn þ − rp þ − ln ;
previously,155–157,160,161,180–182 and should correspond to the jej σ n þ σ p 2kT 2 σn þ σp 2 σ n þ σ p 4 m*n
rhombohedral phase, r8 (Si-XII), and the body-centred cubic (1)
phase, bc8 (Si-III), (Fig. 4(d) and (e)).78,159–161 The
discontinuities in the pressure evolution of some
where e is the electron charge, k is Boltzmann's constant, T is
wavenumbers of these spectra suggested a phase transition  
between them about 3 GPa (the inset in Fig. 12(c)), in the temperature, rn(rp) and m*n m*p are the scattering
agreement with previous studies for Si.155–159 A spectrum we parameters and the effective masses of density of states of
collected after the complete pressure release (Fig. 11(c)) was electrons (holes), respectively.
nearly identical to the one we obtained in the sample It was established that the fundamental indirect band gap
recovered from the electronic-transport studies up to 17 GPa of silicon moderately narrows with pressure,80,168,190 whereas
(Fig. 10(c)). This also agreed with the spectra of Si-III the fundamental gaps of Ge and Ge-dominant alloys, SixGe1−x
polymorphs reported in the literature.155–157,160,161,180–182 (x < 0.2), in contrast, insignificantly increase with pressure
These in situ Raman studies also detected a peak at 288 cm−1 in their cubic-diamond phases.168,169,191 The direct band gaps
(at normal pressure) and followed its pressure evolution on of all the Si–Ge semiconductors were found to widen with
the decompression cycle (Fig. 11(c)). Recall here that we pressure.192–196 Thus, the dramatic pressure-driven changes
observed a similar peak at 294 cm−1 in the sample recovered in the thermopower of the Si–Ge samples we observed (Fig. 5,
after the electronic-transport experiments up to 17 GPa 6, and 9) could not be explained by these pressure
(Fig. 10(c)). Thus, the repeat observation of this peak at a bit dependencies of the band gaps. Hence, in accordance with
different wavenumber in another sample of Si0.974Ge0.026 eqn (1), they should be linked to some intricate changes in
suggested that this is an intrinsic feature, which could be a the electronic band structure parameters, which led to
local mode related to the presence of the Ge atoms in the Si- variations in the ratio of the partial conductivities of
III structure. electrons and holes (σn/σp). Under normal conditions the
Pressure dependencies of the wavenumbers of Seebeck coefficient of the Ge-dominant alloys was found to
Si0.974Ge0.026 for pressurization and decompression cycles for decrease with the Si content (Fig. 5), whereas the
Si-I, Si-II, Si-XII, and Si-III phases are given in Fig. 12. The fundamental band gap should widen with the Si content
main LO phonon of the cubic-diamond Si-I phase shifted to (Fig. 1(c)), thereby suggesting the opposite trend for the
higher frequencies with pressure (Fig. 12(a)). This behaviour thermopower (eqn (1)). Rough estimations for these Ge-
reflected a tightening of the chemical bond in Si-I upon dominant alloys by using eqn (1) with the band-gap values
compression and generally agreed well with previous estimated from Fig. 1(c), k/|e| ≈ 86.4 μV K−1, and 2kT ≈ 50
studies.79,134,150,157,188 However, in contrast to previous meV (at 300 K), gave the partial conductivities ratios as σn/σp
studies for pure Si we found a bend at 2 GPa in the = 0.56 for Si0.02Ge0.98 with S ≈ +330 μV K−1, σn/σp = 0.65 for
pressure dependence of this LO phonon in Si0.974Ge0.026 Si0.07Ge0.93 with S ≈ +272 μV K−1, and σn/σp = 0.77 for
(Fig. 12(a)). This peculiarity had a correspondence with the Si0.13Ge0.87 with S ≈ +182.7 μV K−1 (Fig. 7). In the case of pure
electronic transition to the compensated state we observed germanium the n–p inversion under compression (Fig. 5(a))
in Si0.974Ge0.026 about 2 GPa (Fig. 9). Another peak at 430 was attributed to splitting of overlapping bands of “light”
cm−1, which was assigned to the 2TA(Σ) silicon mode,146 and “heavy” holes on the top of the valence band followed by
showed a similar behaviour with a kink at about 2 GPa a carrier transfer to the ‘light’ one, in which the holes are
(Fig. 12(b)). Meanwhile, a peak at 300 cm−1, which was much more mobile.73
assigned to the 2TA(X) silicon mode,146 demonstrated, in Since the atomic radii of Si and Ge are notably different
contrast, a gradual softening with pressure (Fig. 12(c)). (110 vs. 125 pm) (Fig. 1(a) and (b)), fabrication of their ideal
Thus, our Raman studies of Si0.974Ge0.026 established a solid solutions seems to be not an easy task.197–200 In real Si–
correspondence between the changes in the electronic Ge materials, even in single crystals, a distribution of Si and
transport properties (Fig. 9) on the one hand, and the Ge atoms may not be truly uniform, and a slight tendency to
changes in the vibrational properties (Fig. 11 and 12), which Si/Ge segregation can occur. For example, synchrotron
are linked to the local order, on the other hand. topographic studies of bulk Si–Ge crystals with a low Ge
content (1.4–2.6 at%) observed weak but numerous Ge
striations and showed that high-temperature annealing can
4. Discussion make these crystals more uniform.37,199 Minor negative
deviations of the lattice parameter a of the Si–Ge alloys with
4.1. High-pressure effect on the thermopower of the Si–Ge cubic-diamond structures from the linear Vegard's law were
alloys noted both in epitaxial films201 and in single crystals.37 These
The thermoelectric power of intrinsic semiconductors with findings together with others suggest the presence of local
two-band electrical conductivity depends directly on the band strains in Si–Ge materials, which are linked to variations in
gap value (Eg) and competition between partial conductivities both bond lengths and bond angles between Si(Ge) atoms.202
of electrons (σn) and holes (σp) as follows:189 These strains can affect the properties of the Si–Ge alloys.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5427
Paper CrystEngComm

For example, as found in Raman spectroscopic studies, the acceptor-like levels. However, as seen from their
phonon modes of the Si–Ge alloys with nominally the same thermopower data for the cubic-diamond phase (Fig. 5 and
chemical compositions can have different wavenumbers 6), these acceptor-like levels were very robust, and the
because of the contribution of these spontaneous local samples conserved the p-type conductivity despite the
strains.203,204 apparent irreversible changes in their electronic properties
Thus, for the Si–Ge alloys with intrinsic semiconductor resulting from the defect structure modification. As also seen
conductivity and low values of carrier concentrations, the from these thermopower data (Fig. 5 and 6), the irreversible
pressure-driven changes in their electronic transport changes occurred only on the first pressure cycle, and from
properties should depend not only on variations in their the second cycle the pressure-induced variations were
electronic band structures but also on modifications in their reversible. The irreversible changes in the electronic
native defect structures. For example, the abrupt pressure- properties of the samples after their quenching from high
induced p–n inversion in the conductivity type of the pressure to ambient conditions could be, to some extent,
Si0.974Ge0.026 crystals (Fig. 9) should be linked to the applied- linked to residual strains.
pressure effect on the defect complexes. Germanium atoms
are larger than the silicon ones (Fig. 1(a) and (b)), and they
could be effective traps for vacancies.102,205–208 The enthalpy 4.2. Potential innovative applications of the Si–Ge alloys
of electron ionization in such Ge-vacancy-related complexes The highly spectacular pressure-driven variations in the
in Si rich Si1–xGex crystals was established to increase strongly Seebeck coefficients in the Si–Ge alloys we revealed in this
with the increase in the average Si–Si distance in the cubic- work (Fig. 5–10) hint that these and associated effects can
diamond crystal structure.102,205 From this result it follows have industrial potential. In practical applications, the
that under applied pressure that leads to contraction in the applied-pressure effect can be simulated by various methods,
average Si–Si distance, the enthalpy of electron ionization is which are easier to implement. They include, for example,
expected to reduce progressively, and hence, n-type partial thin epitaxial films deposited on appropriate substrates, in
conductivity can be enhanced. The pressure dependencies of which high internal strains corresponding to pressures of up
the Seebeck coefficient of the Si0.974Ge0.026 crystals up to to 20 GPa can be created,210,211 and nanoindentation with a
about 1 GPa (Fig. 9) were in line with this conjecture. In hard tip, which can generate high stresses in microscopic
addition, the thermopower curves demonstrated that above areas at the sample surface.212,213 Stress-controlled switches
some threshold pressure of about 1 GPa this p–n switching are already utilized in a number of technological processes
becomes irreversible, which can be interpreted as an and electronic devices. For example, on the basis of the
irreversible modification of the defect structure under the nanoindentation technique a new concept of data recording
applied pressure. The thermopower data also disclosed the was offered and a novel memory device was developed by
existence of the second threshold pressure of about 2–4 GPa, IBM company.214,215 It was demonstrated on a sample of
which corresponded to a gradual and irreversible transition of silicon that by using the nanoindentation method one can
Si0.974Ge0.026 crystals to the strongly compensated electronic “write” electrically-conductive areas (metastable Si-III phase)
state (i.e., with σn ≈ σp) (Fig. 9). Likely, this transition was on the surface of semiconducting silicon,212 as well as can
caused by another irreversible modification of the defect stabilize another metastable phase, r8 (Si-XII).216 These
structure, for example, by filling the vacancies and metastable phases can persist under ambient conditions as
subsequent annihilation of their-related complexes. Note that long as they are not annealed.217,218 Note here that such ‘pre-
because of the very low semiconductor-like concentrations of compressed’ metastable phases can be prepared also either
these defects their detailed characterization by other methods by high-pressure torsion,219,220 or by other techniques even
like transmission electron microscopy is hardly possible. not using high pressure, e.g., by colloidal synthesis.221
The stabilization of the p-type conductivity in the Ge- Apparently, the pressure-induced changes in the Seebeck
dominant Si–Ge samples (Fig. 5) may also be linked to the coefficient of the Si–Ge alloys are of potential interest for
occurrence of intrinsic point defects.165,166 It was found that thermoelectric applications of these materials.10–21 However,
the formation of significant number of holes is inevitable for it is more interesting that the effects we discovered suggest
the SiGe layer with high Ge concentration, even if dopants that these inexpensive and environment-friendly materials
are carefully excluded.165 For a set of Si1–xGex (0.19 < x < have potential for the use in nanoelectronic devices as
0.77) films grown by molecular beam epitaxy it was ‘smart’ elements whose optoelectronic properties can be
demonstrated that the p-type conductivity is associated with controlled or switched by applied stress. In Fig. 13 we
acceptor-like states that result from intrinsic point present several simple stress-related applications of the Si–
defects.165,166 GeSi/Si heterostructure research found that the Ge alloys, which can be surmised from the thermopower
main source of holes in SiGe layers is clusters of intrinsic data. High stresses on the surface of these alloys can be
point defects, which are formed due to interaction of generated, for example, by electronically-controlled hard tips
dislocations, but not individual dislocations.209 Likely, the (Fig. 13(a)). This technique can be used to ‘write’ zones or
applied high pressure can also modify the defect structure of electrical circuits with different types of electrical
the Ge-dominant Si–Ge samples and somehow affect these conductivities on the surfaces of the Si–Ge alloys. These

5428 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

Fig. 13 Potential innovative applications of Si–Ge alloys. (a) Schematic side view of hard tip for surface indentation. (b) Top view of hypothetical
the Si–Ge substrate, in which, the electrical conductivity types may be inverted by an appropriate applied stress, as displayed in detail in (c)–(e).
This example illustrates “writing” the zones or circuits of different electrical conductivities. (c–e) Depth profiles of zones whose electrical
conductivity was modified by applied stress. (c) An applied stress above 3 GPa can irreversibly turn the conductivity of n-type Ge to p-type
(Fig. 5(a)), thereby creating a p–n junction. An applied stress higher than 14 GPa can fabricate the metastable st12 polymorph (Ge-III) after stress
release (Fig. 7(a) and (b)). This will form also an underlying p-type layer of Ge-I because of the stress distribution under the hard tip, thereby
creating a n–p–n layered structure. (d) In p-type Ge slightly substituted with Si (p-SixGe1−x) one can form similar layers of the metastable st12
polymorph (Ge-III) by an applied stress above 14 GPa, but the electrical conductivity in these layers can have both n- and p-types, depending on
the Si content (Fig. 7). (e) Applied stresses of 1–2 GPa can switch the conductivity of p-type Si slightly substituted with Ge (p-Si1−xGex) to n-type;
for stresses less than 1 GPa this p–n inversion is reversible, but for stresses above 1.5 GPa it becomes irreversible (Fig. 9(b)). Applied stresses of 2–4
GPa can change p-Si1−xGex into a strongly compensated state with nearly equal partial conductivities of holes and electrons (shown as n ≈ p)
(Fig. 9(a)). Higher applied stresses exceeding 14 GPa allow the formation of the metastable Si-III polymorph with p-type conductivity after stress
release (Fig. 10). This treatment will also result in the formation of two intermediate layers with the compensated (p ≈ n) and n-type conductivity
between the p-Si-III layer and p-Si1−xGex substrate.

‘written’ zones look rather simple from the top (Fig. 13(b)), The thermopower data for the p-type Si1−xGex crystals with
but, in fact, because of the stress distribution under the a low Ge content disclosed more options for the electrical
applied hard tip the depth profiles of these zones may be conductivity variation depending on the applied stress value
intricate. For example, we found that pure Ge with n-type (Fig. 9 and 10). For example, a minor applied stress of about
intrinsic electrical conductivity irreversibly changes into a 1–2 GPa inverted the conductivity to n-type, whereas higher
p-type semiconductor under applied pressure above 3 GPa applied stresses irreversibly changed Si0.974Ge0.026 into the
(Fig. 5(a) and 13(c)). Higher stresses somewhat above the strongly compensated state (i.e., with nearly equal p- and
semiconductor–metal phase transition point76,77,83–92 can n-type partial conductivities) (Fig. 9(a) and 13(e)). Much
stabilize the metastable st12 phase (Ge-III) with the n-type higher applied stresses above the semiconductor–metal
conductivity after stress release (Fig. 7(a) and 13(c)). This phase transition point76–80,86–88,93–96,155–157 resulted in the
treatment should also lead to the formation of an formation of the metastable bc8 phase (Si-III) with a p-type
intermediate p-type layer with the original cubic-diamond conductivity after stress release (Fig. 10(b)). However, two
structure because of the stress distribution in the sample intermediate layers, one with the compensated and the other
(Fig. 13(c)). In germanium crystals with a moderate Si with the n-type conductivity, should be formed here also
content the p-type conductivity became dominant in the because of the stress distribution (Fig. 9 and 13(e)). The
original phase, and for the Si content somewhat exceeding above cases are only the simplest options on how to
7 at.% the p-type conductivity was conserved in the implement the high-pressure effects in Si1−xGex alloys
metastable st12 phase too (Ge-III) (Fig. 7(d)). For the lower revealed in our work. In real applications, this approach may
Si content this metastable phase had an n-type conductivity be extended to fabricate much more intricate cascades of
as in pure Ge (Fig. 7(b) and (c) and 13(d)). diodes, transistors, and entire integrated circuits.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5429
Paper CrystEngComm

Conclusions 5 D. M. Zhigunov, A. B. Evlyukhin, A. S. Shalin, U. Zywietz


and B. N. Chichkov, ACS Photonics, 2018, 5, 977–983.
We synthesized several Si- and Ge-dominant Si–Ge crystals 6 P. R. Abel, A. M. Chockla, Y.-M. Lin, V. C. Holmberg, J. T.
(Si0.974Ge0.026, Si0.02Ge0.98, Si0.07Ge0.93, Si0.13Ge0.87, and Ge) Harris, B. A. Korgel, A. Heller and C. B. Mullins, ACS Nano,
and investigated the applied high-pressure effect on their 2013, 7, 2249–2257.
thermoelectric and electrical properties in the pressure range 7 D. Duveau, B. Fraisse, F. Cunin and L. Monconduit, Chem.
up to 17 GPa at room temperature under multiple pressure Mater., 2015, 27, 3226–3233.
cycling. In addition, the crystal structure of the samples was 8 Y. Yang, S. Liu, X. Bian, J. Feng, Y. An and C. Yuan, ACS
examined by Raman spectroscopy and X-ray diffraction. We Nano, 2018, 12, 2900–2908.
found that the thermoelectric power value of these Si–Ge 9 H. Kim, Y. Son, C. Park, M.-J. Lee, M. Hong, J. Kim, M.
samples in their original cubic-diamond-type phases can be Lee, J. Cho and H. C. Choi, Nano Lett., 2015, 15,
crucially changed by the applied pressure. In some cases even 4135–4142.
the sign inversion of the Seebeck coefficient occurred under 10 J. P. Dismukes, L. Ekstrom, E. F. Steigmeier, I. Kudman and
the applied pressure. Across the phase transitions on D. S. Beers, J. Appl. Phys., 1964, 35, 2899–2907.
pressurization and decompression cycles the thermoelectric 11 O. Yamashita and N. Sadatomi, J. Appl. Phys., 2000, 88,
power of the Si–Ge crystals was significantly varied too. We 245–251.
found that after high-pressure cycling the Si-dominant 12 G. Joshi, H. Lee, Y. Lan, X. Wang, G. Zhu, D. Wang, R. W.
samples transformed into a cubic bc8 phase (Si-III) with a Gould, D. C. Cuff, M. Y. Tang, M. S. Dresselhaus, G. Chen
p-type electrical conductivity, whereas the Ge-dominant alloys and Z. Ren, Nano Lett., 2008, 8, 4670–4674.
changed into a tetragonal st12 phase (Ge-III) whose 13 X. W. Wang, H. Lee, Y. C. Lan, G. H. Zhu, G. Joshi, D. Z.
conductivity type depended on the Si content. Thus, our work Wang, J. Yang, A. J. Muto, M. Y. Tang, J. Klatsky, S. Song,
demonstrated how one can control the thermopower of Si–Ge M. S. Dresselhaus, G. Chen and Z. F. Ren, Appl. Phys. Lett.,
alloys. The unusual pressure behaviour of the Seebeck 2008, 93, 193121.
coefficient of Si–Ge crystals we disclosed points out that they 14 C. Bera, M. Soulier, C. Navone, G. Roux, J. Simon, S. Volz
may be promising for diverse industrial applications, in and N. Mingo, J. Appl. Phys., 2010, 108, 124306.
which the optoelectronic properties of elements are 15 B. Yu, M. Zebarjadi, H. Wang, K. Lukas, H. Wang, D. Wang,
controlled or switched by tuning the applied stress. C. Opeil, M. Dresselhaus, G. Chen and Z. Ren, Nano Lett.,
2012, 12, 2077–2082.
Conflicts of interest 16 E. K. Lee, L. Yin, Y. Lee, J. W. Lee, S. J. Lee, J. Lee, S. N.
Cha, D. Whang, G. S. Hwang, K. Hippalgaonkar, A.
There are no conflicts to declare.
Majumdar, C. Yu, B. L. Choi, J. M. Kim and K. Kim, Nano
Lett., 2012, 12, 2918–2923.
Acknowledgements 17 E. M. J. Hedegaard, S. Johnsen, L. Bjerg, K. A. Borup and
This research was carried out within the state assignment of B. B. Iversen, Chem. Mater., 2014, 26, 4992–4997.
the Ministry of Science and Higher Education of the Russian 18 Y. Lee, A. J. Pak and G. S. Hwang, Phys. Chem. Chem. Phys.,
Federation (theme “Electron” No. AAAA-A18-118020190098-5). 2016, 18, 19544–19548.
The authors thank Dr. A. V. Kurnosov (BGI) for gas loading of 19 R. Murugasami, P. Vivekanandhan, S. Kumaran, R. Suresh
the high-pressure cell, and Prof. L. Dubrovinsky (BGI) for Kumar and T. John Tharakan, Scr. Mater., 2018, 143, 35–39.
providing the high-pressure cell for the Raman spectroscopic 20 K. Delime-Codrin, M. Omprakash, S. Ghodke, R. Sobota, M.
experiments. S. V. O. acknowledges funding from the Adachi, M. Kiyama, T. Matsuura, Y. Yamamoto, M. Matsunami
Deutsche Forschungsgemeinschaft (DFG, German Research and T. Takeuchi, Appl. Phys. Express, 2019, 12, 045507.
Foundation) - Grant No. OV-110/3-1. 21 S. Ahmad, A. Singh, R. Basu, S. Vitta, K. P. Muthe, S. C.
Gadkari and S. K. Gupta, J. Electron. Mater., 2019, 48,
Notes and references 649–655.
22 D. Y. Kim, S. Stefanoski, O. O. Kurakevych and T. A. Strobel,
1 T. Wood, M. Naffouti, J. Berthelot, T. David, J.-B. Claude, L. Nat. Mater., 2015, 14, 169–173.
Métayer, A. Delobbe, L. Favre, A. Ronda, I. Berbezier, N. 23 L. Rapp, B. Haberl, C. J. Pickard, J. E. Bradby, E. G. Gamaly,
Bonod and M. Abbarchi, ACS Photonics, 2017, 4, 873–883. J. S. Williams and A. V. Rode, Nat. Commun., 2015, 6, 7555.
2 S. Sant, S. Lodha, U. Ganguly, S. Mahapatra, F. O. Heinz, L. 24 K. Luo, Z. Zhao, M. Ma, S. Zhang, X. Yuan, G. Gao, X.-F.
Smith, V. Moroz and S. Ganguly, J. Appl. Phys., 2013, 113, Zhou, J. He, D. Yu, Z. Liu, B. Xu and Y. Tian, Chem. Mater.,
033708. 2016, 28, 6441–6445.
3 S. Pouch, M. Amato, M. Bertocchi, S. Ossicini, N. Chevalier, 25 E. M. T. Fadaly, A. Dijkstra, J. R. Suckert, D. Ziss, M. A. J. van
T. Mélin, J.-M. Hartmann, O. Renault, V. Delaye, D. Mariolle Tilburg, C. Mao, Y. Ren, V. T. van Lange, K. Korzun, S.
and Ł. Borowik, J. Phys. Chem. C, 2015, 119, 26776–26782. Kölling, M. A. Verheijen, D. Busse, C. Rödl, J. Furthmüller, F.
4 H. I. T. Hauge, S. Conesa-Boj, M. A. Verheijen, S. Koelling Bechstedt, J. Stangl, J. J. Finley, S. Botti, J. E. M. Haverkort
and E. P. A. M. Bakkers, Nano Lett., 2017, 17, 85–90. and E. P. A. M. Bakkers, Nature, 2020, 580, 205–209.

5430 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

26 Y. Guo, Q. Wang, Y. Kawazoe and P. Jena, Sci. Rep., 2015, 5, 53 J. Schilz and V. N. Romanenko, J. Mater. Sci.: Mater.
14342. Electron., 1995, 6, 265–279.
27 Q. Fan, C. Chai, Q. Wei and Y. Yang, Phys. Chem. Chem. 54 D. Yang, Phys. Status Solidi A, 2005, 202, 931–938.
Phys., 2016, 18, 12905–12913. 55 D. R. Yang and J. Chen, Defect Diffus. Forum, 2005, 242–244,
28 M. Hu, Z. Wang, Y. Xu, J. Liang, J. Li and X. Zhu, Phys. 169–184.
Chem. Chem. Phys., 2018, 20, 26091–26097. 56 D. Yang, X. Yu, X. Ma, J. Xu, L. Li and D. Que, J. Cryst.
29 B. Haberl, T. A. Strobel and J. E. Bradby, Appl. Phys. Rev., Growth, 2002, 243, 371–374.
2016, 3, 040808. 57 A. Misiuk, D. Yang, B. Surma, C. A. Londos, J. Bak-Misiuk
30 M. Amsler, S. Botti, M. A. L. Marques, T. J. Lenosky and S. and A. Andrianakis, Mater. Sci. Semicond. Process., 2006, 9,
Goedecker, Phys. Rev. B: Condens. Matter Mater. Phys., 82–87.
2015, 92, 014101. 58 W. Jung, A. Misiuk and D. Yang, Nucl. Instrum. Methods
31 B. E. Smith, X. Zhou, P. B. Roder, E. H. Abramson and P. J. Phys. Res., 2006, 253, 214–216.
Pauzauskie, J. Appl. Phys., 2016, 119, 185902. 59 S.-L. Sihto, J. Slotte, J. Lento, K. Saarinen, E. V. Monakhov,
32 Z. Ma, X. Liu, X. Yu, C. Shi and F. Yan, Materials, 2017, 10, A. Y. Kuznetsov and B. G. Svensson, Phys. Rev. B: Condens.
599. Matter Mater. Phys., 2003, 68, 115307.
33 Q. Fan, H. Wang, W. Zhang, M. Wei, Y. Song, W. Zhang and 60 A. Chroneos, H. Bracht, C. Jiang, B. P. Uberuaga and R. W.
S. Yun, Curr. Appl. Phys., 2019, 19, 1325–1333. Grimes, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78,
34 J. P. Dismukes, L. Ekstrom and R. J. Paff, J. Phys. Chem., 195201.
1964, 68, 3021–3027. 61 P. Venezuela, G. M. Dalpian, A. J. R. da Silva and A. Fazzio,
35 G. Queisser, W. A. Grosshans and W. B. Holzapfel, EPL, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65, 193306.
1987, 3, 1109–1112. 62 S. N. Guin and K. Biswas, Phys. Chem. Chem. Phys.,
36 A. Werner, J. A. Sanjurjo and M. Cardona, Solid State 2015, 17, 10316–10325.
Commun., 1982, 44, 155–158. 63 T. Nilges, S. Lange, M. Bawohl, J. M. Deckwart, M. Janssen,
37 A. Misiuk, N. V. Abrosimov, P. Romanowski, J. Bak-Misiuk, H.-D. Wiemhöfer, R. Decourt, B. Chevalier, J. Vannahme, H.
A. Wnuk, B. Surma, W. Wierzchowski, K. Wieteska, W. Eckert and R. Weihrich, Nat. Mater., 2009, 8, 101–108.
Graeff and M. Prujszczyk, Mater. Sci. Eng., B, 2008, 154–155, 64 C. Xiao, X. Qin, J. Zhang, R. An, J. Xu, K. Li, B. Cao, J. Yang,
137–140. B. Ye and Y. Xie, J. Am. Chem. Soc., 2012, 134, 18460–18466.
38 R. W. Olesinski and G. J. Abbaschian, Bull. Alloy Phase 65 S. N. Guin, J. Pan, A. Bhowmik, D. Sanyal, U. V. Waghmare
Diagrams, 1984, 5, 283–285. and K. Biswas, J. Am. Chem. Soc., 2014, 136, 12712–12720.
39 M. H. F. Sluiter and Y. Kawazoe, Mater. Trans., 2001, 42, 66 O. Osters, M. Bawohl, J.-L. Bobet, B. Chevalier, R. Decourt
2201–2205. and T. Nilges, Solid State Sci., 2011, 13, 944–947.
40 T. Maeda, H. Hattori, W. H. Chang, Y. Arai and K. 67 P. Roy and T. Maiti, J. Phys. D: Appl. Phys., 2018, 51, 065104.
Kinoshita, Appl. Phys. Lett., 2015, 107, 152104. 68 A. E. Karkin, V. I. Voronin, N. V. Morozova, S. V.
41 S. R. Mehrotra, A. Paul and G. Klimeck, Appl. Phys. Lett., Ovsyannikov, K. Takarabe, Y. Mori, S. Nakamura and V. V.
2011, 98, 173503. Shchennikov, J. Phys. Chem. C, 2016, 120, 9692–9701.
42 A. Levitas, Phys. Rev., 1955, 99, 1810–1814. 69 N. V. Morozova, I. V. Korobeinikov and S. V. Ovsyannikov,
43 R. Braunstein, A. R. Moore and F. Herman, Phys. Rev., J. Appl. Phys., 2019, 125, 220901.
1958, 109, 695–710. 70 V. V. Shchennikov, N. P. Gavaleshko and F. M. Frasunyak,
44 E. R. Johnson and S. M. Christian, Phys. Rev., 1954, 95, Phys. Solid State, 1995, 37, 1943–1944.
560–561. 71 S. V. Ovsyannikov, I. V. Korobeinikov, N. V. Morozova, A.
45 A. Levitas, C. C. Wang and B. H. Alexander, Phys. Rev., Misiuk, N. V. Abrosimov and V. V. Shchennikov, Appl. Phys.
1954, 95, 846–846. Lett., 2012, 101, 062107.
46 J. Weber and M. I. Alonso, Phys. Rev. B: Condens. Matter 72 S. V. Ovsyannikov, A. E. Karkin, N. V. Morozova, V. V.
Mater. Phys., 1989, 40, 5683–5693. Shchennikov, E. Bykova, A. M. Abakumov, A. A. Tsirlin,
47 F. Herman, Phys. Rev., 1954, 95, 847–848. K. V. Glazyrin and L. Dubrovinsky, Adv. Mater., 2014, 26,
48 T. Honda, M. Suezawa and K. Sumino, Jpn. J. Appl. Phys., 8185–8191.
1996, 35, 5980–5985. 73 I. V. Korobeinikov, N. V. Morozova, V. V. Shchennikov and
49 X. Guo, Z. Tian, T. Gao, Q. Xie, Y. Liang, Y. Mo and Y. S. V. Ovsyannikov, Sci. Rep., 2017, 7, 44220.
Wanjun, Phys. Chem. Chem. Phys., 2017, 19, 4695–4700. 74 N. V. Morozova, I. V. Korobeinikov, K. V. Kurochka, A. N.
50 V. Saltas, A. Chroneos and F. Vallianatos, Sci. Rep., 2017, 7, Titov and S. V. Ovsyannikov, J. Phys. Chem. C, 2018, 122,
1374. 14362–14372.
51 S.-R. G. Christopoulos, N. Kuganathan and A. Chroneos, 75 T. Wen, Y. Wang, N. Li, Q. Zhang, Y. Zhao, W. Yang, Y. Zhao
Sci. Rep., 2019, 9, 10849. and H.-K. Mao, J. Am. Chem. Soc., 2019, 141, 505–510.
52 V. G. Deibuk, S. I. Shakhovtsova, V. A. Shenderovski and 76 S. Minomura and H. G. Drickamer, J. Phys. Chem. Solids,
V. M. Tsmots, Semicond. Phys., Quantum Electron. 1962, 23, 451–456.
Optoelectron., 2002, 5, 5–8. 77 J. C. Jamieson, Science, 1963, 139, 762–764.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5431
Paper CrystEngComm

78 R. H. Wentorf and J. S. Kasper, Science, 1963, 139, 338–339. 103 D. J. Frost, B. T. Poe, R. G. Trønnes, C. Liebske, A. Duba
79 B. A. Weinstein and G. J. Piermarini, Phys. Rev. B: Solid and D. C. Rubie, Phys. Earth Planet. Inter., 2004, 143–144,
State, 1975, 12, 1172–1186. 507–514.
80 B. Welber, C. K. Kim, M. Cardona and S. Rodriguez, Solid 104 S. V. Ovsyannikov, H. Gou, A. E. Karkin, V. V. Shchennikov,
State Commun., 1975, 17, 1021–1024. R. Wirth, V. Dmitriev, Y. Nakajima, N. Dubrovinskaia and
81 D. Olego and M. Cardona, Phys. Rev. B: Condens. Matter L. S. Dubrovinsky, Chem. Mater., 2014, 26, 5274–5281.
Mater. Phys., 1982, 25, 1151–1160. 105 S. V. Ovsyannikov, M. Bykov, E. Bykova, D. P. Kozlenko,
82 T. Soma, H. Iwanami and H. Matsuo, Solid State Commun., A. A. Tsirlin, A. E. Karkin, V. V. Shchennikov, S. E.
1982, 42, 469–471. Kichanov, H. Gou, A. M. Abakumov, R. Egoavil, J. Verbeeck,
83 M. Baublitz and A. L. Ruoff, J. Appl. Phys., 1982, 53, C. McCammon, V. Dyadkin, D. Chernyshov, S. van Smaalen
5669–5671. and L. S. Dubrovinsky, Nat. Chem., 2016, 8, 501–508.
84 S. B. Qadri, E. F. Skelton and A. W. Webb, J. Appl. Phys., 106 S. V. Ovsyannikov, M. Bykov, E. Bykova, K. Glazyrin, R. S.
1983, 54, 3609–3611. Manna, A. A. Tsirlin, V. Cerantola, I. Kupenko, A. V.
85 C. S. Menoni, J. Z. Hu and I. L. Spain, Phys. Rev. B: Condens. Kurnosov, I. Kantor, A. S. Pakhomova, I. Chuvashova, A. I.
Matter Mater. Phys., 1986, 34, 362–368. Chumakov, R. Rüffer, C. McCammon and L. S.
86 H. Olijnyk, S. K. Sikka and W. B. Holzapfel, Phys. Lett. A, Dubrovinsky, Nat. Commun., 2018, 9, 4142.
1984, 103, 137–140. 107 S. V. Ovsyannikov, E. Bykova, A. Pakhomova, D. P.
87 H. Olijnyk, Phys. Rev. Lett., 1992, 68, 2232–2234. Kozlenko, M. Bykov, S. E. Kichanov, N. V. Morozova, I. V.
88 A. Mujica, A. Rubio, A. Muñoz and R. J. Needs, Rev. Mod. Korobeinikov, F. Wilhelm, A. Rogalev, A. A. Tsirlin, A. V.
Phys., 2003, 75, 863–912. Kurnosov, Y. G. Zainulin, N. I. Kadyrova, A. P. Tyutyunnik
89 A. Di Cicco, A. C. Frasini, M. Minicucci, E. Principi, J.-P. Itiè and L. Dubrovinsky, Inorg. Chem., 2017, 56, 6251–6263.
and P. Munsch, Phys. Status Solidi B, 2003, 240, 19–28. 108 N. V. Morozova, S. V. Ovsyannikov, I. V. Korobeinikov, A. E.
90 A. Di Cicco, E. Principi, M. Minicucci, S. De Panfilis, A. Karkin, K. Takarabe, Y. Mori, S. Nakamura and V. V.
Filipponi, F. Decremps, F. Datchi, J.-P. Itié, P. Munsch and Shchennikov, J. Appl. Phys., 2014, 115, 213705.
A. Polian, High Pressure Res., 2004, 24, 93–99. 109 I. Kantor, V. Prakapenka, A. Kantor, P. Dera, A. Kurnosov, S.
91 E. Principi, A. Di Cicco, F. Decremps, A. Polian, S. De Sinogeikin, N. Dubrovinskaia and L. Dubrovinsky, Rev. Sci.
Panfilis and A. Filipponi, Phys. Rev. B: Condens. Matter Instrum., 2012, 83, 125102.
Mater. Phys., 2004, 69, 201201. 110 A. Kurnosov, I. Kantor, T. Boffa-Ballaran, S. Lindhardt, L.
92 F. Coppari, J. C. Chervin, A. Congeduti, M. Lazzeri, A. Dubrovinsky, A. Kuznetsov and B. H. Zehnder, Rev. Sci.
Polian, E. Principi and A. Di Cicco, Phys. Rev. B: Condens. Instrum., 2008, 79, 045110.
Matter Mater. Phys., 2009, 80, 115213. 111 S. Klotz, J.-C. Chervin, P. Munsch and G. Le Marchand,
93 J. Z. Hu, L. D. Merkle, C. S. Menoni and I. L. Spain, Phys. J. Phys. D: Appl. Phys., 2009, 42, 075413.
Rev. B: Condens. Matter Mater. Phys., 1986, 34, 4679–4684. 112 V. V. Shchennikov, S. V. Ovsyannikov and A. Y. Manakov,
94 G. A. Voronin, C. Pantea, T. W. Zerda, L. Wang and Y. Zhao, J. Phys. Chem. Solids, 2010, 71, 1168–1174.
Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 020102. 113 V. V. Shchennikov, S. V. Ovsyannikov and A. V. Bazhenov,
95 H. Katzke, U. Bismayer and P. Tolédano, Phys. Rev. B: J. Phys. Chem. Solids, 2008, 69, 2315–2324.
Condens. Matter Mater. Phys., 2006, 73, 134105. 114 V. V. Shchennikov, N. V. Morozova and S. V. Ovsyannikov,
96 J. J. Guo, D. Pan, X. Q. Yan, T. Fujita and M. W. Chen, Appl. J. Appl. Phys., 2012, 111, 112624.
Phys. Lett., 2010, 96, 251910. 115 F. P. Bundy, Phys. Rep., 1988, 167, 133–176.
97 S. Anzellini, M. T. Wharmby, F. Miozzi, A. Kleppe, D. 116 L. G. Khvostantsev, V. N. Slesarev and V. V. Brazhkin, High
Daisenberger and H. Wilhelm, Sci. Rep., 2019, 9, 15537. Pressure Res., 2004, 24, 371–383.
98 F. Coppari, A. Polian, N. Menguy, A. Trapananti, A. 117 A. K. Bandyopadhyay, S. Chatterjee, E. S. R. Gopal and S. V.
Congeduti, M. Newville, V. B. Prakapenka, Y. Choi, E. Subramanyam, Rev. Sci. Instrum., 1981, 52, 1232–1235.
Principi and A. Di Cicco, Phys. Rev. B: Condens. Matter 118 I. V. Korobeinikov, N. V. Morozova, L. N. Lukyanova, O. A.
Mater. Phys., 2012, 85, 045201. Usov, V. A. KulbachinskII, V. V. Shchennikov and S. V.
99 F. Coppari, A. Di Cicco, A. Congeduti, J. C. Chervin, F. Ovsyannikov, J. Phys. D: Appl. Phys., 2018, 51, 025501.
Baudelet and A. Polian, High Pressure Res., 2009, 29, 119 M. I. Eremets, High Pressure Experimental Methods, Oxford
103–107. University Press, Oxford–New York–Toronto, 1996.
100 F. Coppari, A. Di Cicco, E. Principi, A. Trapananti, N. Pinto, 120 A. Celeste, F. Borondics and F. Capitani, High Pressure Res.,
A. Polian, S. Chagnot and A. Congeduti, High Pressure Res., 2019, 39, 608–618.
2010, 30, 28–34. 121 T. Pippinger, R. Miletich, M. Merlini, P. Lotti, P.
101 N. V. Abrosimov, S. N. Rossolenko, W. Thieme, A. Schouwink, T. Yagi, W. A. Crichton and M. Hanfland, Phys.
Gerhardt and W. Schröder, J. Cryst. Growth, 1997, 174, Chem. Miner., 2015, 42, 29–43.
182–186. 122 V. V. Shchennikov, S. V. Ovsyannikov, A. Y. Derevskov and
102 V. P. Markevich, A. R. Peaker, L. I. Murin and N. V. V. V. Shchennikov, J. Phys. Chem. Solids, 2006, 67,
Abrosimov, Appl. Phys. Lett., 2003, 82, 2652–2654. 2203–2209.

5432 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

123 S. V. Ovsyannikov, X. Wu, G. Garbarino, M. Núñez- 146 O. Moutanabbir, S. Miyamoto, E. E. Haller and K. M. Itoh,
Regueiro, V. V. Shchennikov, J. A. Khmeleva, A. E. Karkin, Phys. Rev. Lett., 2010, 105, 026101.
N. Dubrovinskaia and L. Dubrovinsky, Phys. Rev. B: 147 B. C. Johnson, B. Haberl, J. E. Bradby, J. C. McCallum and
Condens. Matter Mater. Phys., 2013, 88, 184106. J. S. Williams, Phys. Rev. B: Condens. Matter Mater. Phys.,
124 S. V. Ovsyannikov, N. V. Morozova, I. V. Korobeinikov, L. N. 2011, 83, 235205.
Lukyanova, A. Y. Manakov, A. Y. Likhacheva, A. I. Ancharov, 148 F. P. Bundy and J. S. Kasper, Science, 1963, 139, 340–341.
A. P. Vokhmyanin, I. F. Berger, O. A. Usov, V. A. Kutasov, 149 J. M. Besson, E. H. Mokhtari, J. Gonzalez and G. Weill,
V. A. KulbachinskII, T. Okada and V. V. Shchennikov, Appl. Phys. Rev. Lett., 1987, 59, 473–476.
Phys. Lett., 2015, 106, 143901. 150 T. P. Mernagh and L.-G. Liu, J. Phys. Chem. Solids, 1991, 52,
125 S. V. Ovsyannikov, N. V. Morozova, I. V. Korobeinikov, V. 507–512.
Haborets, R. Yevych, Y. VysochanskII and V. V. 151 M. Oh-ishi, S. Akiyama, K. Uchida, S. Nozaki and H.
Shchennikov, Dalton Trans., 2017, 46, 4245–4258. Morisaki, Phys. Status Solidi B, 2001, 223, 391–395.
126 N. V. Morozova, V. V. Shchennikov and S. V. Ovsyannikov, 152 A. Mujica and R. J. Needs, Phys. Rev. B: Condens. Matter
J. Appl. Phys., 2015, 118, 225901. Mater. Phys., 1993, 48, 17010–17017.
127 D. W. Feldman, M. Ashkin and J. H. Parker, Phys. Rev. Lett., 153 A. Mujica, S. Radescu, A. Munoz and R. J. Needs, Phys.
1966, 17, 1209–1212. Status Solidi B, 2001, 223, 379–384.
128 W. J. Brya, Solid State Commun., 1973, 12, 253–257. 154 R. J. Nelmes, M. I. McMahon, N. G. Wright, D. R. Allan and
129 J. S. Lannin, Phys. Rev. B: Solid State, 1977, 16, 1510–1518. J. S. Loveday, Phys. Rev. B: Condens. Matter Mater. Phys.,
130 H. D. Fuchs, C. H. Grein, M. I. Alonso and M. Cardona, 1993, 48, 9883–9886.
Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 44, 155 M. Hanfland and K. Syassen, High Pressure Res., 1990, 3,
13120–13123. 242–244.
131 Z. Sui, H. H. Burke and I. P. Herman, Phys. Rev. B: Condens. 156 H. Olijnyk and A. P. Jephcoat, Phys. Status Solidi B,
Matter Mater. Phys., 1993, 48, 2162–2168. 1999, 211, 413–420.
132 F. Pezzoli, L. Martinelli, E. Grilli, M. Guzzi, S. Sanguinetti, 157 L. Q. Huston, A. Lugstein, J. S. Williams and J. E. Bradby,
M. Bollani, H. D. Chrastina, G. Isella, H. von Känel, E. Appl. Phys. Lett., 2018, 113, 123103.
Wintersberger, J. Stangl and G. Bauer, Mater. Sci. Eng., B, 158 J. Crain, G. J. Ackland, J. R. Maclean, R. O. Piltz, P. D.
2005, 124–125, 127–131. Hatton and G. S. Pawley, Phys. Rev. B: Condens. Matter
133 D. Rouchon, M. Mermoux, F. Bertin and J. M. Hartmann, Mater. Phys., 1994, 50, 13043–13046.
J. Cryst. Growth, 2014, 392, 66–73. 159 R. O. Piltz, J. R. Maclean, S. J. Clark, G. J. Ackland, P. D.
134 T. Ishidate, S. Katagiri, K. Inoue, M. Shibuya, K. Tsuji and Hatton and J. Crain, Phys. Rev. B: Condens. Matter Mater.
S. Minomura, J. Phys. Soc. Jpn., 1984, 53, 2584–2591. Phys., 1995, 52, 4072–4085.
135 C. H. Grein and M. Cardona, Phys. Rev. B: Condens. Matter 160 O. O. Kurakevych, Y. Le Godec, W. A. Crichton, J. Guignard,
Mater. Phys., 1992, 45, 8328–8333. T. A. Strobel, H. Zhang, H. Liu, C. Coelho Diogo, A. Polian,
136 M. Franz, K. F. Dombrowski, H. Rücker, B. Dietrich, K. N. Menguy, S. J. Juhl and C. Gervais, Inorg. Chem., 2016, 55,
Pressel, A. Barz, U. Kerat, P. Dold and K. W. Benz, Phys. Rev. 8943–8950.
B: Condens. Matter Mater. Phys., 1999, 59, 10614–10621. 161 Z. Zeng, Q. Zeng, W. L. Mao and S. Qu, J. Appl. Phys.,
137 S. Rath, M. L. Hsieh, P. Etchegoin and R. A. Stradling, 2014, 115, 103514.
Semicond. Sci. Technol., 2003, 18, 566–575. 162 J.-T. Wang, C. Chen, H. Mizuseki and Y. Kawazoe, Phys. Rev.
138 O. Pagès, J. Souhabi, V. J. B. Torres, A. V. Postnikov and Lett., 2013, 110, 165503.
K. C. Rustagi, Phys. Rev. B: Condens. Matter Mater. Phys., 163 B. Haberl, M. Guthrie, B. D. Malone, J. S. Smith, S. V.
2012, 86, 045201. Sinogeikin, M. L. Cohen, J. S. Williams, G. Shen and J. E.
139 K. Gao, S. Prucnal, A. Mücklich, W. Skorupa and S. Zhou, Bradby, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89,
Acta Phys. Pol., A, 2013, 123, 858–861. 144111.
140 M. I. Alonso and K. Winer, Phys. Rev. B: Condens. Matter 164 Z. Zhao, H. Zhang, D. Y. Kim, W. Hu, E. S. Bullock and T. A.
Mater. Phys., 1989, 39, 10056–10062. Strobel, Nat. Commun., 2017, 8, 13909.
141 A. Picco, E. Bonera, E. Grilli, M. Guzzi, M. Giarola, G. 165 M. Satoh, K. Arimoto, K. Nakagawa, S. Koh, K. Sawano, Y.
Mariotto, D. Chrastina and G. Isella, Phys. Rev. B: Condens. Shiraki, N. Usami and K. Nakajima, Jpn. J. Appl. Phys.,
Matter Mater. Phys., 2010, 82, 115317. 2008, 47, 4630–4633.
142 J. S. Reparaz, A. Bernardi, A. R. Goñi, M. I. Alonso and M. 166 M. Satoh, K. Arimoto, J. Yamanaka, K. Nakagawa, K.
Garriga, Appl. Phys. Lett., 2008, 92, 081909. Sawano and Y. Shiraki, Jpn. J. Appl. Phys., 2012, 51,
143 J. B. Renucci, R. N. Tyte and M. Cardona, Phys. Rev. B: Solid 105801.
State, 1975, 11, 3885–3895. 167 W. Paul and D. M. Warschauer, J. Phys. Chem. Solids,
144 X. S. Zhao, Y. R. Ge, J. Schroeder and P. D. Persans, Appl. 1958, 5, 89–101.
Phys. Lett., 1994, 65, 2033–2035. 168 W. Paul, J. Phys. Chem. Solids, 1959, 8, 196–204.
145 P. Mishra and K. P. Jain, Phys. Rev. B: Condens. Matter 169 K. Tanaka, Phys. Rev. B: Condens. Matter Mater. Phys.,
Mater. Phys., 2001, 64, 073304. 1991, 43, 4302–4307.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5433
Paper CrystEngComm

170 R. J. Kobliska, S. A. Solin, M. Selders, R. K. Chang, R. 198 M. Arivanandhan, Y. Saito, T. Koyama, Y. Momose, H.
Alben, M. F. Thorpe and D. Weaire, Phys. Rev. Lett., Ikeda, A. Tanaka, T. Tatsuoka, D. K. Aswal, Y. Inatomi and
1972, 29, 725–728. Y. Hayakawa, J. Cryst. Growth, 2011, 318, 324–327.
171 S. Nozaki, S. Sato, S. Rath, H. Ono and H. Morisaki, Bull. 199 N. V. Abrosimov, A. Lüdge, H. Riemann and W. Schröder,
Mater. Sci., 1999, 22, 377–381. J. Cryst. Growth, 2002, 237-239, 356–360.
172 S. Rath, S. Sato, H. Ono, S. Nozaki and H. Morisaki, Mater. 200 A. I. Veinger, A. G. Zabrodski, T. V. Tisnek, S. I.
Chem. Phys., 1998, 54, 244–246. Goloshchapov and N. V. Abrosimov, Semiconductors,
173 A. Kailer, K. G. Nickel and Y. G. Gogotsi, J. Raman 2007, 41, 666–672.
Spectrosc., 1999, 30, 939–946. 201 E. Kasper, A. Schuh, G. Bauer, B. Holländer and H. Kibbel,
174 B. D. Malone and M. L. Cohen, Phys. Rev. B: Condens. J. Cryst. Growth, 1995, 157, 68–72.
Matter Mater. Phys., 2012, 86, 054101. 202 I. Yonenaga and M. Sakurai, Phys. Rev. B: Condens. Matter
175 L. Q. Huston, B. C. Johnson, B. Haberl, S. Wong, J. S. Mater. Phys., 2001, 64, 113206.
Williams and J. E. Bradby, J. Appl. Phys., 2017, 122, 175108. 203 F. Pezzoli, E. Bonera, E. Grilli, M. Guzzi, S.
176 Y. Ikoma, K. Kumano, K. Edalati, K. Saito, Q. Guo and Z. Sanguinetti, D. Chrastina, G. Isella, H. von Känel, E.
Horita, Philos. Mag. Lett., 2017, 97, 27–34. Wintersberger, J. Stangl and G. Bauer, J. Appl. Phys.,
177 D. J. Oliver, J. E. Bradby, J. S. Williams, M. V. Swain and P. 2008, 103, 093521.
Munroe, J. Appl. Phys., 2007, 101, 043524. 204 M. R. Islam, M. Yamada, N. V. Abrosimov, M. Kiyama
178 G. Andersson, B. Sundqvist and G. Bäckström, J. Appl. and M. Tatsumi, Eur. Phys. J.: Appl. Phys., 2004, 27,
Phys., 1989, 65, 3943–3950. 325–328.
179 S. V. Ovsyannikov, V. V. Shchennikov, A. Misiuk and V. V. 205 V. P. Markevich, A. R. Peaker, J. Coutinho, R. Jones, V. J. B.
Shchennikov, Solid State Commun., 2004, 132, 545–549. Torres, S. Öberg, P. R. Briddon, L. I. Murin, L. Dobaczewski
180 G. Weill, J. L. Mansot, G. Sagon, C. Carlone and J. M. and N. V. Abrosimov, Phys. Rev. B: Condens. Matter Mater.
Besson, Semicond. Sci. Technol., 1989, 4, 280–282. Phys., 2004, 69, 125218.
181 A. Kailer, Y. G. Gogotsi and K. G. Nickel, J. Appl. Phys., 206 A. R. Peaker, V. P. Markevich, F. D. Auret, L. Dobaczewski
1997, 81, 3057–3063. and N. Abrosimov, J. Phys.: Condens. Matter, 2005, 17,
182 M. M. Khayyat, G. K. Banini, D. G. Hasko and M. M. S2293–S2302.
Chaudhri, J. Phys. D: Appl. Phys., 2003, 36, 1300–1307. 207 A. R. Peaker, V. P. Markevich, B. Hamilton, I. D. Hawkins,
183 A. Wosylus, H. Rosner, W. Schnelle and U. Schwarz, Z. J. Slotte, K. Kuitunen, F. Tuomisto, A. Satta, E. Simoen
Anorg. Allg. Chem., 2009, 635, 700–703. and N. V. Abrosimov, Thin Solid Films, 2008, 517,
184 H. Zhang, H. Liu, K. Wei, O. O. Kurakevych, Y. Le Godec, Z. 152–154.
Liu, J. Martin, M. Guerrette, G. S. Nolas and T. A. Strobel, 208 J. Chen, T. Wu, X. Ma, L. Wang and D. Yang, J. Appl. Phys.,
Phys. Rev. Lett., 2017, 118, 146601. 2008, 103, 123519.
185 H. Aharoni, Vacuum, 1978, 28, 571–578. 209 P. N. Grillot, S. A. Ringel, J. Michel and E. A. Fitzgerald,
186 M. F. A. Alias, N. Rammo and M. Makadsi, Renewable J. Appl. Phys., 1996, 80, 2823–2832.
Energy, 2001, 24, 347–351. 210 M. Baleva and E. Mateeva, Phys. Rev. B: Condens. Matter
187 C. Xu, C. L. Senaratne, R. J. Culbertson, J. Kouvetakis and J. Mater. Phys., 1994, 50, 8893–8896.
Menéndez, J. Appl. Phys., 2017, 122, 125702. 211 A. Kadir, T. Ganguli, R. Kumar, M. R. Gokhale, A. P. Shah,
188 N. N. Ovsyuk and S. G. Lyapin, Appl. Phys. Lett., 2020, 116, S. Ghosh, B. M. Arora and A. Bhattacharya, Appl. Phys. Lett.,
062103. 2007, 91, 111913.
189 K. Seeger, Semiconductor Physics, Springer, New York, 1973. 212 S. Ruffell, K. Sears, J. E. Bradby and J. S. Williams, Appl.
190 W. Paul and D. M. Warschauer, J. Phys. Chem. Solids, Phys. Lett., 2011, 98, 052105.
1958, 5, 102–106. 213 S. Wong, B. Haberl, J. S. Williams and J. E. Bradby, Appl.
191 W. Paul and D. M. Warschauer, J. Phys. Chem. Solids, Phys. Lett., 2015, 106, 252103.
1958, 6, 6–15. 214 P. Vettiger, G. Cross, M. Despont, U. Drechsler, U. Durig, B.
192 M. Cardona and W. Paul, J. Phys. Chem. Solids, 1960, 17, Gotsmann, W. Haberle, M. A. Lantz, H. E. Rothuizen, R.
138–142. Stutz and G. K. Binnig, IEEE Trans. Nanotechnol., 2002, 1,
193 B. Welber, M. Cardona, Y.-F. Tsay and B. Bendow, Phys. Rev. 39–55.
B: Solid State, 1977, 15, 875–879. 215 M. I. Lutwyche, M. Despont, U. Drechsler, U. Dürig, W.
194 A. R. Goni, K. Syassen and M. Cardona, Phys. Rev. B: Häberle, H. Rothuizen, R. Stutz, R. Widmer, G. K.
Condens. Matter Mater. Phys., 1989, 39, 12921–12924. Binnig and P. Vettiger, Appl. Phys. Lett., 2000, 77,
195 G. G. N. Angilella, N. H. March, I. A. Howard and R. Pucci, 3299–3301.
J. Phys.: Conf. Ser., 2008, 121, 032006. 216 S. Wong, B. Haberl, B. C. Johnson, A. Mujica, M. Guthrie,
196 C. V. de Alvarez and M. L. Cohen, Solid State Commun., J. C. McCallum, J. S. Williams and J. E. Bradby, Phys. Rev.
1974, 14, 317–320. Lett., 2019, 122, 105701.
197 R. H. Deitch, S. H. Jones and T. G. Digges, J. Electron. 217 S. Mannepalli and K. S. R. N. Mangalampalli, J. Appl. Phys.,
Mater., 2000, 29, 1074–1078. 2019, 125, 225105.

5434 | CrystEngComm, 2020, 22, 5416–5435 This journal is © The Royal Society of Chemistry 2020
CrystEngComm Paper

218 S. Wong, B. C. Johnson, B. Haberl, A. Mujica, J. C. 220 B. Chon, Y. Ikoma, M. Kohno, J. Shiomi, M. R.
McCallum, J. S. Williams and J. E. Bradby, J. Appl. Phys., McCartney, D. J. Smith and Z. Horita, Scr. Mater.,
2019, 126, 105901. 2018, 157, 120–123.
219 Y. Ikoma, B. Chon, T. Yamasaki, K. Takahashi, K. Saito, Q. 221 S. Ganguly, N. Kazem, D. Carter and S. M. Kauzlarich,
Guo and Z. Horita, Appl. Phys. Lett., 2018, 113, 101904. J. Am. Chem. Soc., 2014, 136, 1296–1299.

This journal is © The Royal Society of Chemistry 2020 CrystEngComm, 2020, 22, 5416–5435 | 5435

View publication stats

You might also like