ArticlePDF Available

Abstract and Figures

Abstract Strain relaxation processes in InAs heteroepitaxy have been studied. While InAs grows in a layer-by-layer mode on lattice-mismatched substrates of GaAs(111)A, Si(111), and GaSb(111)A, the strain relaxation process strongly depends on the lattice mismatch. The density of threading defects in the InAs film increases with lattice mismatch. We found that the peak width in x-ray diffraction is insensitive to the defect density, but critically depends on the residual lattice strain in InAs films.
This content is subject to copyright. Terms and conditions apply.
1
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
Strain relaxation in InAs
heteroepitaxy on lattice-
mismatched substrates
Akihiro Ohtake*, Takaaki Mano & Yoshiki Sakuma
Strain relaxation processes in InAs heteroepitaxy have been studied. While InAs grows in a layer-by-
layer mode on lattice-mismatched substrates of GaAs(111)A, Si(111), and GaSb(111)A, the strain
relaxation process strongly depends on the lattice mismatch. The density of threading defects in the
InAs lm increases with lattice mismatch. We found that the peak width in x-ray diraction is insensitive
to the defect density, but critically depends on the residual lattice strain in InAs lms.
Heteroepitaxy of semiconductors has opened up new possibilities for band-structure engineering and novel
devices, including strained-layer structures. e exibility in the choice of materials for the formation of het-
erostructures is oen limited by lattice mismatch. Heteroepitaxy in lattice-mismatched systems usually follows
a Stranski-Krastanov (SK) growth mode: a pseudomorphic two-dimensional layer is formed below a certain
critical thickness, and is followed by the formation of three-dimensional islands. A prototypical example of such
a system is Ge on Si (lattice mismatch 4.2%), in which layer-by-layer growth is limited to 3–4 monolayer (ML).
e island formation is highly undesirable, because it prevents the growth of smooth lms and introduces nucle-
ation centers for defects. us, signicant eorts have been devoted to suppress the strain-induced islanding and
strain-relieving defects. It has been reported that the islanding in the Ge/Si system is eectively suppressed by
introducing As and Sb as surfactant species13.
e SK growth occurs also in the InAs/GaAs(001) system having lattice mismatch of 7.2%: InAs islands are
formed at the lm thickness of 1.6 ML4. On the other hand, the use of the (111)A-oriented GaAs substrates forces
the InAs lm to grow in a layer-by-layer mode57. e layer-by-layer growth of the (111)A-oriented InAs lm is
accompanied by the formation of a mist dislocation network at the InAs/GaAs interface6, so that the generation
of defects in the lm is strongly suppressed. Similar strain relaxation has been reported for GaSb/GaAs(001) het-
eroepitaxy, which is known for so-called interfacial mist array growth8.
e novel growth technique has been successfully applied to the layer-by-layer growth of InAs on Si(111)9,
GaSb growth on InAs/Si(111)10, and to the improvement of the crystalline quality of InGaAs11 and GaSb12
on InAs/GaAs(111)A. is technique has a great advantage, especially for the growth on Si(111), because the
formation of antiphase domain boundaries in InAs films is suppressed, in contrast with the growth on the
(001)-oriented substrate. However, strain relaxation processes of InAs are far from being completely understood,
and the crystalline quality of InAs has not been studied in detail.
is paper reports the strain relaxation processes and structural properties of InAs lms heteroepitaxially
grown on the lattice-mismatched substrates of GaAs(111)A, Si(111), and GaSb(111)A. While the InAs(111)A
lm grows in a layer-by-layer mode on all substrates irrespective of the lattice mismatch, the strain relaxation
process behaves dierently depending on the lattice mismatch. Fully-strained pseudomorphic InAs layers con-
tinue to grow above ~50 ML on the nearly lattice matched GaSb substrates (lattice mismatch of 0.61%), while
in the InAs/Si system with the largest lattice mismatch of 11.4%, mostly-relaxed InAs lms are formed even at
the very initial stage of the growth. e in-plane compressive strain in InAs on GaAs (lattice mismatch of 7.2%)
gradually relaxed as the growth proceeds, but the strain is not fully relaxed even in the 100nm-thick InAs lm.
Our x-ray diraction (XRD) measurements revealed that the residual strain in InAs lm is responsible for the
peak broadening in XRD proles. On the other hand, the peak width of x-ray rocking curve is insensitive to the
density of threading defects.
National Institute for Materials Science (NIMS), Tsukuba, 305-0044, Japan. *email: OHTAKE.Akihiro@nims.go.jp
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
Results and discussion
 Figure1(a) shows the variation in the in-plane lattice constant (d110) of the InAs lm growing on the GaAs(111)
A substrate. e d110 values were measured from the distance between the 11 and
1
1
reections in reection
high-energy electron diraction (RHEED) patterns along the [
1
1
2] direction. e data clearly shows that InAs
pseudomorphically grows below ~2 ML, and that the in-plane lattice constant gradually increases with lm thick-
ness above ~2 ML, in good agreement with earlier results6,7. e strain in the InAs lm has relaxed by only ~80%,
even aer the 50 ML-growth.
Similarly to the case for InAs/GaAs(111)A, the InAs lm is two-dimensionally grown on the In-terminated
Si(111) substrate9. However, the strain relaxation processes between the two systems are quite dierent. Shown
in Fig.1(b) is the variation in the d110 value for the InAs growth on the Si(111) substrate. A new set of streaks
from the InAs lm appeared in the RHEED patterns at the very early stage of the growth (~0.6 ML), in addition
to those from the Si substrate. e spacing of streaks is quite close to the value of bulk InAs, and remains almost
unchanged throughout the growth (<50 ML). is indicates that, on the Si(111) substrate, the InAs lm was
nucleated with its inherent lattice constant, and that pseudomorhic InAs layers are not formed. us, it is likely
that the lattice mismatch of InAs/Si (11.5%) is too large to be accommodated by elastic deformation of thin InAs
lms.
In the nearly lattice-matched system of InAs on GaSb(111)A (lattice mismatch 0.61%), as shown in Fig.1(c),
no signicant change in the d110 value is observed below 50 ML (17.5 nm). us, it is plausible that 50 ML-InAs
lms are coherently strained to the GaSb substrate. is is consistent with the critical thickness (~20 nm) for
the onset of mist dislocations in InAs on GaSb estimated on the basis of the model proposed by Matthews and
Blakeslee13.
Figure 1. e variation of the in-plane lattice constant (d110) of InAs lms grown on the GaAs(111)A (a),
Si(111) (b), and GaSb(111)A (c) substrates. e values were measured from the distance between the 11 and
1
1
reections in the RHEED patterns.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
Figure2(a,b) show the two-dimensional reciprocal-space maps (RSMs) for the asymmetric 115 reection of
100 nm-thick InAs lms grown on the GaAs(111)A and In-terminated Si(111) substrates, respectively. e verti-
cal and horizontal axes correspond to the indices along the [111] and [
1
1
2] directions, respectively. e in-plane
lattice constant of InAs grown on GaAs is 0.4236 nm, and is quite smaller than that of bulk InAs (0.4284 nm). e
residual strain in the InAs lm (d110 (lm)d110(bulk))/d110(bulk)) is 1.124%. According to classical elastic
theory, the in-plane compressive strain causes the expansion of the lattice constant, d111, in the direction normal
to the surface. Poissons ratio σ of the strained InAs(111)A layer is σ=(c11+2c12+4c44)/
(2c11+4c124c44)=1.75, where c11=8.33,c12=4.53, and c44=3.96 (×106 Pa) are the elastic stiness coe-
cients of InAs14,15. us, using the measured d110 value, the d111 value is estimated to be 0.3520 nm, in good agree-
ment with the measured value of 0.3511 nm.
e d110 and d111 values of InAs grown on Si(111) are 0.4281 and 0.3499 nm, which are quite close to those
of bulk InAs (0.4284 nm and 0.3498 nm), as seen in Fig.2(b). The residual in-plane strain is estimated to
be 0.089%, indicating that the InAs lm is mostly relaxed.
In contrast to the case for the compressively strained systems of InAs/GaAs and InAs/Si, InAs lms are tensile
strained on GaSb. e measured d110 (0.4303 nm) and d111 (0.3489 nm) values of the InAs lm are slightly larger
and smaller, respectively, than those of bulk values. e tensile in-plane strain in InAs is relaxed by ~30%, indicat-
ing that the strain begins to relax in the InAs lm below 100 nm. is is broadly consistent with the mechanical
equilibrium model13, from which the critical thickness is estimated to be 20 nm, as mentioned earlier. On the other
hand, the growth experiments on the (001)-oriented GaSb substrate showed that InAs lms as thick as 200 nm are
fully strained16. It is suggested that the strain relaxation mechanism is a strong function of substrate orientation.
Figure3(a–c) compare x-ray rocking curves (XRCs) measured from 100nm-InAs on GaAs(111)A, Si(111),
and GaSb(111)A. e FWHM values are 1627.7 arcsec (a), 230.05 arcsec (b), and 496.83 arcsec (c). It is well
known that crystal imperfections, such as mosaic domains, threading defects, structural inhomogeneities, and
lattice distortions, cause a broadening of XRD proles. However, it is dicult to specify the origin of the peak
broadening from the rocking curves of the symmetric 111 reection alone. On the other hand, RSMs of asymmet-
ric reections provides important information of the structural quality of the InAs lm; as can be seen in Fig.3(a),
the 115 reection has an elliptical shape, which is elongated along the direction perpendicular to the [115] azi-
muth. is means that the broadening of XRD peaks arises from the misorientation of InAs lattice planes. While a
slight broadening is observed in Fig.2(c), such a broadening is not observed in Fig.2(b). In Fig.3(d), the FWHM
values are plotted as a function of the absolute value of residual strain: the FWHM value increases with the abso-
lute value of residual strain. In general, uniformly-strained at layers are unstable against the modulation of the
surface prole, which allows a partial relaxation of the strain by elastic deformation1719. us, it is likely that the
strained InAs lm on the GaAs(111)A substrate is elastically deformed to accommodate the residual strain. Such
elastic deformation is accompanied by the misorientation of the InAs lattice planes, resulting in the broadening
perpendicular to the reciprocal lattice vector of the corresponding diraction spot.
Figure4(a) shows a representative plan-view transmission electron microscopy (TEM) image of InAs on
Si(111). A high density of threading defects, such as threading dislocations, stacking faults, and stacking-fault
tetrahedra, is clearly observed. e densitiesof these defects are listed in Table1. Similar TEM images were
obtained for InAs on GaAs(111)A and GaSb(111)A with lower densities of threading defects. On the other hand,
Figure 2. Reciprocal-space maps (RSMs) of the asymmetric 115 reection measured from 100 nm-InAs lms
on GaAs (a), Si (b), and GaSb (c) substrates. e dashed lines show the position of bulk InAs. e RSMs were
generated using the Bruker DIFFRAC.LEPTOS soware package ver. 7.7, and were assembled using the Adobe
Illustrator CS6.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
as mentioned earlier, the FWHM value of XRC for InAs/Si is much smaller than that for InAs/GaAs. e FWHM
value of XRC is oen cited as a measure of structural quality: it has been generally believed that the existence of
high densities of threading defects causes the broadening of XRC width. However, the present results show that
narrower (broader) peaks in XRCs are not necessarily an indication of lower (higher) density of threading defects
in heteroepitaxial layers.
Previous studies have shown that the two-dimensional growth of InAs on GaAs(111)A and Si(111) substrates
is accompanied by the formation of a mist dislocation network at the interfaces to accommodate the strain6,7. In
the ideal case, perfect dislocation arrays are formed covering all the interface area, leaving the growing lms free
of threading dislocations. However, as shown in Fig.4(b), the actual mist dislocation arrays are not perfect; it is
likely that the disorder in the mist dislocation network acts as the source of the threading defects (arrows). e
averaged period of mist-dislocation network pmd is given by pmd=b/f, where b is the Burgers vector and f is the
lattice mist. For partially-relaxed 100 nm-InAs on GaAs(111)A, the pmd value is estimated to be 6.7 nm, which
is twice as large as that for the fully-relaxed interface of InAs/Si(111) (3.3 nm). If we assume that the nucleation
probability of threading defects increases with the density of mist dislocations, we can explain why the defect
density is higher in the InAs lm on Si(111).
In the nearly lattice-matched InAs/GaSb system, the densities of threading defects are slightly lower than
those for InAs/GaAs, as shown in Table1. On the other hand, since the lattice mismatch of InAs/GaSb is more
than an order of magnitude smaller than that for InAs/GaAs, one may expect a much lower density of mist dis-
locations. Figure4(c) shows the plan-view TEM image of 100nm-InAs/GaSb(111)A; since the sample contains
both the InAs lm and the GaSb substrate, mist dislocations at the interface are imaged (yellow dashed lines), as
well as threading defects (red arrows). We note that only threading defects in the InAs lms are imaged at thinner
(<100 nm) regions (See Supplementary Fig. S1). As compared with InAs on GaAs (Fig.4(b)), the distribution
of the mist dislocations is highly irregular; such an irregular conguration is likely to be responsible for the
increase of threading defects: as indicated by arrows in Fig.4(c), threading dislocations are oen observed at
the end of mist dislocations. us, it is reasonable to consider that the extremely large critical thickness for the
generation of mist dislocation makes it dicult to rearrage highly separated mist dislocations to an ordered
periodic network at the buried interface beneath thick InAs layers.
Figure 3. 111 XRCs of 100 nm-InAs lms grown on GaAs(111)A (a), Si(111) (b), and GaSb(111)A (c).
(d) FWHM values of the 111 XRCs plotted as a function of residual strain.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
Conclusions
e eects of lattice mismatch on the growth mode and strain relaxation process in InAs heteroepitaxy on
GaSb(111)A, GaAs(111)A, and Si(111) substrates have been studied. e highest and lowest defect densities are
identied in the highly lattice-mismatched InAs/Si and the nearly lattice-matched InAs/GaSb systems, respec-
tively. On the other hand, the peak width in XRD is insensitive to the lattice mismatch and is roughly proportional
to the residual strain in InAs lms.
Figure 4. Plan-view TEM images of InAs lms grown on Si(111) (a) and GaSb(111)A (c) substrates. (b) shows
the scanning tunneling microscopy (STM) image of 4 ML-InAs on GaAs(111)A. Image dimensions of (a–c) are
900 nm×1200 nm, 200 nm×150 nm, and 1500 nm×2000 nm, respectively. e STM image was acquired in
the constant current mode with a tunneling current of 0.1 nA and a sample voltage of 3 V. e arrows marked
TD, SF, and SFT, indicate the threading dislocation, stacking fault, and stacking-fault tetrahedron, respectively,
while the position of the mist dislocation (MD) is indicated by the yellow dashed lines. We note that the Si
substrate was completely removed from the sample (a), while the sample (c) consists of the InAs lm and GaSb
substrate.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
Methods
e growth experiments were carried out in a multi-chamber MBE system. e As-doped Si(111) wafer was
cleaned by radiatively heating at 950°C in the MBE chamber20. Clean and well-ordered (7×7) reconstructions
were conrmed by STM, x-ray photoelectron spectroscopy, and RHEED. In-terminated (4×1) reconstruction
was prepared by depositing 1 ML of In on the Si(111)-(7×7) surface at 450°C. e clean surfaces of GaAs(111)
A and GaSb(111)A were prepared by growing undoped homoepitaxial layers at 450°C on the thermally cleaned
substrates21,22. e GaAs (GaSb) layers were grown with an As4/Ga (Sb4/Ga) ux ratio of ~50 (~8). While only a
(2×2) reconstruction is observed on the GaAs(111)A surface under the conventional MBE condition, (2×2)
(2
3
×23
) and (1×5) reconstructions were observed on GaSb(111)A as the surface Sb coverage is increased. In
the present study, GaAs(111)A-(2×2) and GaSb(111)A-(2
3
23×
) surfaces were used as substrates for the InAs
growth.
InAs lms were grown on the GaAs(111)A substrate at 450 °C with an As4/In ux ratio of ~50. On the other
hand, the growth on Si(111) and GaSb(111)A results in the formation of twins and islands under otherwise iden-
tical condition. us, to suppress the growth front roughening, the growth on these substrates were carried out
using As2 molecules having higher reactivity to In adatoms with a higher As2/In ratio of ~150. e growth rate of
InAs was approximately 0.034 ML/s, which was calibrated by RHEED intensity oscillation measurements on the
(001)-oriented InAs substrate. Here, 1 ML of InAs is dened as 6.3×1014 atoms/cm2, which is the site-number
density of unreconstructed InAs(111)A surface. Great care was taken to avoid the possible adsorption of As
molecules on Si(111) substrate for the growth on the In-terminated Si(111) substrate, because the In-terminated
Si surface easily reacts with As molecules to transform itself to the As-terminated one, on which InAs grows in
an island mode9. us, the InAs growth on Si is initiated under the lower As2/In ratio of 30, and the ux ratio is
increased to ~150 aer the 10 ML-growth. e growth process of InAs was monitored by RHEED in real time,
and the structural properties have been characterized using XRD and TEM. High resolution XRD measurements
were carried out using a monochromatic Cu Kα1 radiation. A channel-cut analyzer crystal was used for XRC
measurements. RSM data were obtained using an one-dimensional array detector. e density and type of thread-
ing defects in InAs lms were assessed by TEM operated at 200 keV.
Data availability
e data that support the ndings of this study are available from the corresponding author upon reasonable
request.
Received: 25 November 2019; Accepted: 31 January 2020;
Published: xx xx xxxx
References
1. Copel, M., euter, M. C., axiras, E. & Tromp, . M. Surfactants in epitaxial growth. Phys. Rev. Lett. 63, 632–635 (1989).
2. Copel, M., euter, M. C., Horn-vonHoegen, M. & Tromp, . M. Inuence of surfactants in Ge and Si epitaxy on Si(001). Phys. Rev.
B 42, 11682–11689 (1990).
3. Horn-vonHoegen, M., LeGoues, F. ., Copel, M., euter, M. C. & Tromp, . M. Defect self-annihilation in surfactant-mediated
epitaxial growth. Phys. Rev. Lett. 67, 1130–1133 (1991).
4. Leonard, D., Pond, . & Petro, P. M. Critical layer thicness for self-assembled InAs islands on GaAs. Phys. Rev. B 50, 11687–11692
(1994).
5. Yamaguchi, H., Fahy, M. . & Joyce, B. A. Inhibitions of three dimensional island formation in InAs lms grown on GaAs(111)A
surface by molecular beam epitaxy. Appl. Phys. Lett. 69, 776–778 (1996).
6. Yamaguchi, H. et al. Atomic-scale imaging of strain relaxation via misfit dislocations in highly mismatched semiconductor
heteroepitaxy: InAs/GaAs(111)A. Phys. Rev. B 55, 1337–1340 (1997).
7. Ohtae, A., Ozei, M. & Naamura, J. Strain relaxation in InAs/GaAs(111)A heteroepitaxy. Phys. Rev. Lett. 84, 4665–4668 (2000).
8. Huang, S. H. et al. Strain relief by periodic mist arrays for low defect density GaSb on GaAs. Appl. Phys. Lett. 88, 131911 (2006).
9. Ohtae, A. & Mitsuishi, . Polarity controlled InAs{111} lms grown on Si(111). J. Vac. Sci. Technol. B 29, 031804 (2011).
10. Ohtae, A., Mano, T., Miyata, N., Mori, T. & Yasuda, T. Heteroepitaxy of GaSb on Si(111) and fabrication of HfO2 /GaSb metal-
oxide-semiconductor capacitors. Appl. Phys. Lett. 104, 032101 (2014).
11. Mano, T. et al. Growth of metamorphic InGaAs on GaAs(111)A: Counteacting lattice mismatch by inserting a thin InAs interlayer.
Cryst. Growth Des. 16, 5412–5417 (2016).
12. Ohtae, A., Mano, T., Mitsuishi, . & Sauma, Y. Strain relaxation in GaSb/GaAs(111)A heteroepitaxy using thin InAs interlayers.
ACS Omega 3, 15592–15597 (2018).
13. Matthews, J. W. & Blaeslee, A. E. Defects in epitaxial multilayers: I. Mist dislocations. J. Cryst. Growth 27, 118–125 (1974).
14. Marttin, . M. Elastic properties of ZnS structure semiconductors. Phys. Rev. B 1, 4005–4011 (1970).
15. Hornstra, J. & Bartels, W. J. Determination of the lattice constant of epitaxial layers of III-V compounds. J. Cryst. Growth 44, 513–517
(1978).
16. Bennett, B. . Strain relaxation in InAs/GaSb heterostructures. Appl. Phys. Lett. 73, 3736–3738 (1998).
17. Srolovitz, D. J. On the stability of surfaces of stressed solids. Acta Metall 37, 621–625 (1989).
18. Piddu, A. J., obbins, D. J., Cullis, A. G., Leong, Y. Y. & Pitt, A. M. Evolution of surface morphology and strain during SiGe epitaxy.
in Solid Films 222, 78–84 (1992).
InAs/GaAs(111)A InAs/Si(111) InAs/GaSb(111)A
threading dislocation 2.7×1097.5×1091.7×109
stacking fault 3.0×1095.1×1091.9×109
stacking-fault tetrahedron 8.2×1083.0×1095.4×108
Table 1. Density of threading defects in InAs on GaAs(111)A, Si(111), and GaSb(111)A [cm2].
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
SCIENTIFIC REPORTS | (2020) 10:4606 | https://doi.org/10.1038/s41598-020-61527-9
www.nature.com/scientificreports
www.nature.com/scientificreports/
19. Spencer, B. J., Voorhees, P. W. & Davis, S. H. Morphological instability in epitaxially strained dislocation-free solid lms. Phys. Rev.
Lett. 67, 3696–3699 (1991).
20. Ohtae, A. Surface reconstructions on GaAs(001). Surf. Sci. Rep. 63, 295 (2008).
21. Ohtae, A., Ha, N. & Mano, T. Extremely High- and low-density of Ga droplets on GaAs{111}A,B: surface-polarity dependence.
Cryst. Growth Des. 15, 485–488 (2015).
22. Miyata, N., Ohtae, A., Ichiawa, M., Mori, T. & Yasuda, T. Electrical characteristics and thermal stability of HfO2 metal-oxide-
semiconductor capacitors fabricated on clean reconstructed GaSb surfaces. Appl. Phys. Lett. 104, 232104 (2014).
Acknowledgements
is work was partlysupported by JSPS KAKENHI Grant Number 19K04480.
Author contributions
A.O. conceived and conducted the growth experiments, T.M. and Y.S. carried out the XRD measurements. All
authors reviewed the manuscript.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41598-020-61527-9.
Correspondence and requests for materials should be addressed to A.O.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2020
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... The residual strain in the InAs is <0.2%. 25 A cross-sectional TEM image of the sample shows MD formation at the interface of InAs/GaAs, which is marked in red circles ( Figure 2c). Most of the MDs are confined at the interface. ...
... As reported previously, the imperfection of the MDs at the interface probably causes the generation of TDs. 22,25 We also found stacking faults (SFs) and stacking fault tetrahedra (SFTs) (Figure 2d) along the <011> direction, which are often seen in the layer grown on (111)A surfaces. 25 The total density of the SFs and SFTs is 7.7 × 10 8 cm −2 . ...
... 22,25 We also found stacking faults (SFs) and stacking fault tetrahedra (SFTs) (Figure 2d) along the <011> direction, which are often seen in the layer grown on (111)A surfaces. 25 The total density of the SFs and SFTs is 7.7 × 10 8 cm −2 . We used a higher growth rate for InAs (1 ML s −1 ) compared to our previous work (0.03 ML s −1 ). 25 However, the defect density (TD, SF, and SFT) is nearly the same (or even less). ...
Article
We demonstrate an extended short-wave infrared (e-SWIR) photodetector composed of an InAs/GaAs(111)A heterostructure with interface misfit dislocations. The layer structure of the photodetector consists simply of an n-InAs optical absorption layer directly grown with a thin undoped-GaAs spacer layer on n-GaAs by molecular beam epitaxy. The lattice mismatch was abruptly relaxed by forming a misfit dislocation network at the initial stage of the InAs growth. We found high-density threading dislocations (1.5 × 109 cm-2) in the InAs layer. The current-voltage characteristics of the photodetector at 77 K had a very low dark current density (<1 × 10-9 A cm-2) at a positive applied voltage (electrons flow from n-GaAs to n-InAs) of up to ∼+1 V. Simulation of the band structure revealed that the direct connection of GaAs and InAs and the formation of interfacial states by the misfit dislocations play significant positive roles in suppressing dark current. Under illumination with e-SWIR light at 77 K, a clear photocurrent signal was observed with a 2.6 μm cutoff wavelength, which is consistent with the bandgap of InAs. We also demonstrated e-SWIR detection at room temperature with a 3.2 μm cutoff wavelength. The maximum detectivity at 294 K exceeds 2 × 108 cm Hz0.5 W-1 for the detection of e-SWIR light at 2 μm.
... 21) Earlier studies have shown that in the lattice-mismatched system of InAs on GaAs(111)A, the formation of three-dimensional islands is effectively inhibited by introducing misfit-dislocation network at the interface. [23][24][25][26] The layer-by-layer growth continues throughout the growth, which is in stark contrast to that for the (100) orientation where Stranki-Krastanow growth occurs. [27][28][29] The layer-by-Page 2 of 12 AUTHOR SUBMITTED MANUSCRIPT -JJAP-S1103907.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 A c c e p t e d M a n u s c r i p t layer growth was also realized for InSb on GaAs(111)A by combining the two-step growth technique, and the 200 nm-thick-InSb layer exhibits mobility close to ~10,000 cm 2 V -1 s -1 . ...
... Thus, it is likely that the lattice mismatch of InSb/GaAs(111)A (14.6 %) is too large to be accommodated by elastic deformation of thin InSb films, similarly to the case for InAs/Si (11.5 %). 25) To reduce the lattice mismatch between InSb and GaAs, the growth experiments were carried out using thin InAs interlayers. Figure 2 Here, it is interesting to note that the strain relaxation proceeds more rapidly for thinner InAs thickness. ...
Article
Full-text available
Molecular-beam epitaxy of InSb on the (111)A-oriented GaAs substrates has been studied using electron diffraction, x-ray diffraction, and scanning probe microscopy. The direct heteroepitaxialgrowth of InSb on GaAs(111)A results in a cracked morphology with flat terraces and deep gaps, which could be attributed to the extremely large lattice mismatch between InSb and GaAs (14.6 %). When thin (5 – 30 monolayer thickness) InAs films are used as interlayers, more continuous and flat InSb films are obtained. The proposed growth technique using (111)A-oriented GaAs substrates and thin InAs interlayers are effective in improving the surface morphology and the structural quality of InSb films in highly lattice-mismatched systems.
... The XRD results reveal sharp and overlapped peaks, proving an abruptness of p-InAs interface with underlayers and uniformity across the wafer. On the other hand, FWHM of a layer peak can be used to qualitatively compare the crystal quality, associated with threading defect density (dislocations), mosaic domain or wafer curvature, as long as the samples share identical structure and are measured using the same XRD instrument and conditions [52]. The average FWHM ω values with standard deviations are 167.4±2.5 and 159.4±2.2 arcsec for (004) and (115) reflections, respectively [ Fig. 5(c)]. ...
... ω (115) is much lower than reported for InAs on (111) GaAs and (111) GaSb [52]. In conclusion, these XRD datasets demonstrate slight broadening across the wafer with the lowest FWHM values at the centre. ...
Article
Full-text available
The article presents the results of diameter mapping for circular-symmetric disturbance of homogeneity of epitaxially grown InAs (100) layers on GaAs substrates. The set of acceptors (beryllium) doped InAs epilayers was studied in order to evaluate the impact of Be doping on the 2-inch InAs-on-GaAs wafers quality. During the initial identification of size and shape of the circular pattern, non-destructive optical techniques were used, showing a 100% difference in average roughness between the wafer centre and its outer part. On the other hand, no volumetric (bulk) differences are detectable using Raman spectroscopy and highresolution X-ray diffraction. The correlation between Be doping level and circular defect pattern surface area has been found.
... Two common ways to induce strain on semiconductor thin films are (1) lattice-mismatched epitaxial growth and (2) external application through the diamond anvil cell (DAC) method [7]. Lattice-mismatched epitaxial growth consists of growing a semiconductor thin film on a substrate with a different lattice constant, resulting in heterostructures like p-doped indium arsenide on gallium antimonide (p-InAs/GaSb) [8] and gallium arsenide on silicon (GaAs/Si) [9]. The type of strain applied to the film depends on how the lattice constant of the film compares with its substrate. ...
... While both methods are widely used in strain-related research, they have their own disadvantages. Lattice-mismatched epitaxial growth is subject to strain relaxation effects which create crystal defects like threading dislocations or cracks [8]. The effects of strain on a lattice-mismatched heterostructure cannot be isolated in strain-related studies because crystal defects also affect carrier dynamics and transport properties like phonon scattering rates [10]. ...
Article
Full-text available
We investigate strain effects on the ultrafast carrier dynamics and transport of gallium arsenide films on silicon (GaAs/Si) and magnesium oxide (GaAs/MgO) substrates using temperature-dependent photoluminescence (PL) and terahertz time-domain spectroscopy (THz-TDS) from 11 K - 300 K. The PL shows that GaAs/Si and GaAs/MgO samples are under tensile and compressive strain at low temperature, respectively. The temperature-dependent THz emission from GaAs/Si does not show significant differences with the emission from bulk GaAs, while the THz emission from GaAs/MgO shows an order-of-magnitude decrease at low temperature. The THz emission from the samples exhibits an interplay between strain-induced effective mass changes and temperature-dependent electric field effects.
... where ν, E are the Poisson's ratio and the Young's modulus of film, respectively; and S is the surface area of film. Figure 1 shows the results of calculating the coordinate dependence of crystal lattice deformation of the InAs film on the GaAs substrate (Figure 1a Table 1 [23][24][25]. ...
Article
Full-text available
In our work, the model of self-consistent electron–deformation–diffusion effects in thin films grown on substrate with the mismatch of lattice parameters of the contacting materials is constructed. The proposed theory self-consistently takes into account the interaction of the elastic field (created by the mismatch of lattice parameters of the film and the substrate, and point defects) with the diffusion processes of point defects and the electron subsystem of semiconductor film. Within the framework of the developed model, the spatial distribution of deformation, concentration of defects, conduction electrons and electric field intensity is investigated, depending on the value of the mismatch, the type of defects, the average concentrations of point defects and conduction electrons. It is established that the coordinate dependence of deformation and the concentration profile of defects of the type of stretching (compression) centers, along the axis of growth of the strained film, have a non-monotonic character with minima (maxima), the positions of which are determined by the average concentration of point defects. It is shown that due to the electron–deformation interaction in film with a lattice parameter mismatch, the spatial redistribution of conduction electrons is observed and n-n+ transitions can occur. Information about the self-consistent spatial redistribution of point defects, electrons and deformation of the crystal lattice in semiconductor materials is necessary for understanding the problems of their stability and degradation of nano-optoelectronic devices operating under conditions of intense irradiation.
... According to the inset of Fig. 1, the peak broadening of the O@MoS2 in respect to the pristine MoS2 attributes to an increase in strain or lattice deformation [43]. Moreover, the partial replacement of oxygen with sulfur is associated with the left-shift of the O@MoS2 peaks due to the increase of the interlayer spacing [44]. ...
Article
Full-text available
High electronic transport and reasonable chemical stability of molybdenum disulfide (MoS2) make it very suitable for electrochemical applications. However, its energy storage capacity is still low compared with other nanostructures. In this work, pristine and thermally oxidized MoS2 ([email protected]2) based hybrids are introduced by a simple method with enhanced capacitive performance thanks to the contribution of synergistic effects. Scanning electron microscopy (SEM), Transmission electron microscopy (TEM), X-ray diffraction (XRD), elemental mapping, UV-Visible, Raman, X-ray photoelectron (XPS) spectroscopies and BET specific surface area analyses are employed to investigate the morphological and crystalline structure of the introduced hybrids. In detail, the highest gravimetric capacitance of ∼205.1 Fg⁻¹ is achieved for the MoS2:[email protected]2 hybrid with a mass ratio of 2:1 compared to pristine and other electrodes. This electrode is also accompanied by the longest discharging time and excellent cyclic stability of ∼%113 after 2000 continuous charge-discharge cycles. In addition, photoelectrochemical testing of the introduced electrode leads to a ∼63% increase in carrier photogeneration compared to MoS2 due to the effective charge separation within the hybrid, which makes it suitable for water splitting and hydrogen production applications.
Chapter
This chapter focuses on providing a systematic analysis of the far-infrared (FIR) reflectivity/transmission, Raman scattering, spectroscopic ellipsometry (SE), and high-resolution synchrotron extended x-ray absorption spectroscopy (HR-SXAS) data for comprehending film thickness, optical, structural, and phonon traits of several epitaxially grown II–VI/III–V epilayers and superlattices (SLs). The structural and electronic properties are acquired by employing phonon Raman scattering, induced by deformation potential, as well as coupled plasmon LO-phonon modes. Experimental FIR line shapes and transverse optical modes (ωTO) have offered information on epilayer alignment to the substrates. A traditional methodology of multilayer optics (ambient/ film/substrate) is exploited for analyzing the FIR reflectivity/transmission spectra at both near normal incidence (θa = 0°) and oblique incidence (θa = 45°), i.e., using the Berreman effect, for extracting film thickness d of binary/ternary epilayers and layered structures of SLs. A clear distinction of the transverse optical (ωTO) in the s-polarization, longitudinal optical (ωLO), and ωTO modes in the p-polarization has established an effective way of assessing the long wavelength optical phonons in ultrathin materials. Model dielectric functions derived from the SE data are meticulously incorporated to fit the reflectivity and transmission spectra of thin binary and ternary alloys in the range of ~1.0 to 6 eV. Comprehensive analyses of the HR-SXAS data on the structural characteristics of epilayers have provided accurate values of the bond lengths and coordination numbers in very good agreement with the existing appropriate data of the bulk materials. Examples of epitaxially grown semiconductor materials considered here included the CdZnTe/GaAs(001), CdTe/InSb(001) epilayers, and heavily strained (ZnSe)m/(BeSe)n/GaAs(001) SLs.
Article
This paper reviews recent developments in the lattice‐mismatched epitaxy of InAs on (111)A‐oriented substrates and related research topics, in which the presence or absence of the misfit dislocations is controlled via prescribed growth sequences. When InAs is grown on GaAs (111)A substrates under standard growth conditions, a unique lattice‐relaxation mechanism occurs. A misfit dislocation network is formed at the initial stage of InAs growth, that is followed by the layer‐by‐layer growth of relaxed InAs films. The InAs/GaAs (111)A heterostructure is being applied in infrared photodetectors which have a new operating principle that employs the high density dislocations at the interface. The InAs/GaAs (111)A heterostructure is also useful for the growth of InGaAs layers: a nearly lattice‐relaxed InGaAs containing different concentrations of indium can be formed by inserting a thin‐InAs layer between the InGaAs and GaAs. These InGaAs layers can be used as virtual substrates with a desired lattice constant for a range of devices. In addition to the formation of lattice‐relaxed structures, dislocation‐free InAs QDs can be formed on InP (111)A substrates by applying droplet epitaxy. The C 3v symmetry of the (111)A surface makes it possible to form symmetric InAs quantum dots that emit entangled photon pairs at telecommunication wavelengths. This article is protected by copyright. All rights reserved.
Article
Full-text available
We have systematically studied the strain relaxation processes in GaSb heteroepitaxy on GaAs(111)A using thin InAs interlayers. The growth with 1 ML- and 2 ML-InAs leads to formation of an InAsSb-like layer, which induces tensile strain in GaSb films, whereas the GaSb films grown with thicker InAs layers (≥3 ML) are under compressive strain. As the InAs thickness is increased above 5 ML, the insertion of the InAs layer becomes less effective in the strain relaxation, leaving residual strain in GaSb films. This leads to the elastic deformation of the GaSb lattice, giving rise to the increase in the peak width of X-ray rocking curves.
Article
We have successfully grown high quality InxGa1-xAs metamorphic layer on GaAs (111)A using molecular beam epitaxy. Inserting a thin 3.0 - 7.1 monolayer (ML) InAs interlayer between the In0.25Ga0.75As and GaAs allowed the formation of a nearly lattice-relaxed In0.25Ga0.75As with a very flat upper surface. However, when the thickness of the inserted InAs is thinner or thicker than these values, we observed degradation of crystal quality and/or surface morphology. We also revealed that this technique to be applicable to the formation of a high quality metamorphic InxGa1-xAs layer with a range of In compositions (0.25≦ x≦ 0.78) on GaAs (111)A. Cross-sectional scanning transmission electron microscope studies revealed that misfit dislocations formed only at the interface of InAs and GaAs, not at the interface of In0.25Ga0.75As and InAs. From the dislocation density analysis, it is suggested that the dislocation density was decreased by growing In0.25Ga0.75As on InAs, which effectively contribute the strain relaxation of In0.25Ga0.75As. The InGaAs/InAlAs quantum wells that were formed on the metamorphic layers exhibit clear photoluminescence emissions up to room temperature.
Article
Formation processes of Ga droplets on polar (111)A and (111)B surfaces of GaAs have been investigated. A single Ga atom forms a stable nucleus on the (111)A surface, so that the formation of extremely high-density of Ga droplets is achieved (2.8 x 10(12) cm(-2)). On the (111)B surface, the initial Ga deposition on both As-rich (2 X 2) and Ga-rich (root 19 x root 19) reconstructions leads to the formation of a two-dimensional GaAs layer having a more Ga-rich (3 x 2) reconstruction. The Ga droplets are formed on the (3 X 2) surface with their densities being 4 orders of magnitude lower than those for the (111)A orientation.
Article
HfO2/GaSb interfaces fabricated by high-vacuum HfO2 deposition on clean reconstructed GaSb surfaces were examined to explore a thermally stable GaSb metal-oxide-semiconductor structure with low interface-state density (Dit). Interface Sb-O bonds were electrically and thermally unstable, and post-metallization annealing at temperatures higher than 200 °C was required to stabilize the HfO2/GaSb interfaces. However, the annealing led to large Dit in the upper-half band gap. We propose that the decomposition products that are associated with elemental Sb atoms act as interface states, since a clear correlation between the Dit and the Sb coverage on the initial GaSb surfaces was observed.
Article
The (111)A-oriented GaSb films are two-dimensionally grown on the Si(111) substrate. We found that the insertion of a thin interface layer of InAs between GaSb and Si is very effective to obtain high-quality GaSb films. Using the GaSb/InAs/Si heterostructure, we have fabricated HfO2/GaSb metal-oxide-semiconductor (MOS) capacitors. The MOS capacitors show electrical characteristics comparable to those fabricated on GaSb(001) substrates, making itself suitable for realizing the integration of Sb-based MOS devices with Si substrates.
Article
A comparison has been made of the surface morphology of thin InAs films grown on GaAs (001) and (111)A substrates by molecular beam epitaxy using in situ reflection high energy electron diffraction and ex situ atomic force microscopy. InAs growth on (001) surface proceeds via the Stranski‐Krastanov mechanism, with three‐dimensional island formation beginning between one and two monolayers, but on the (111)A surface there is a two‐dimensional mode, independent of detailed growth conditions. This advantage accruing from the use of a novel index substrate provides the opportunity of fabricating a wide range of high quality heterostructures.
Article
This paper reviews the recent experimental findings on the atomic structures on the (001) surface of GaAs. We systematically studied the structure and composition of the GaAs(001) surfaces using reflection high-energy electron diffraction, reflectance difference spectroscopy, scanning tunneling microscopy, and X-ray photoelectron spectroscopy. We found that the As-rich c(4×4)β, c(4×4)α, and (2×4), and Ga-rich (6×6), c(8×2), and (4×6) reconstructions are formed on the GaAs(001) surface critically depending on the preparation conditions. Atomic structures on these reconstructions will be discussed on the basis of the recent findings of experiments and first-principles calculations.
Article
When a thin layer of a slightly different lattice constant is grown coherently on a single crystal substrate, the lattice constant of the layer material cannot be measured directly, due to elastic strains. It can be calculated from quantities observed by X-ray double crystal diffractometry if the anisotropic elasticity is taken into account. The way of calculation is shown for arbitrary surface orientation, and the effect of misfit dislocations is indicated.
Article
Strain relaxation in InAs/GaAs(111)A heteroepitaxy has been studied on the atomic scale by scanning tunneling microscopy. The coalescence of small islands and the formation of a dislocation network are identified at the critical layer thickness (CLT), and no three-dimensional growth is observed, even beyond the CLT. The atomic displacement around the threading segments and the strain fields induced by the misfit dislocations are both identified. The measured density of the misfit dislocations indicates that the strain is not fully relaxed at the CLT, but is instead gradually relieved with the additional growth of InAs.
Article
A simple phenomenological theory of the elastic constants of sphalerite structure crystals is presented and shown to apply within reasonable errors to the known experimental constants. The theory utilizes a form for bond-stretching (α) and bending (β) forces first used by Keating, to which are added effective point-ion Coulombic forces. Also it is pointed out that regularities in the experimental elastic constants of these crystals are readily explained in terms of the ionicity fi defined by Phillips and Van Vechten. Of particular note are the shear constants which decrease markedly with ionicity. It is found that this decrease is described quantitatively by β/α∝(1-fi), which confirms the interpretation of β, since bond-bending forces should vanish in the ionic limit fi→1. Other equally simple formulas for the forces in terms of only the bond length and fi are shown to predict all the constants with a rms accuracy of 10%.