ArticlePDF Available

InAs on GaAs Photodetectors Using Thin InAlAs Graded Buffers and Their Application to Exceeding Short-Wave Infrared Imaging at 300 K

Authors:

Abstract and Figures

Short-wave infrared (SWIR) detectors and emitters have a high potential value in several fields of applications, including the internet of things (IoT) and advanced driver assistance systems (ADAS), gas sensing. Indium Gallium Arsenide (InGaAs) photodetectors are widely used in the SWIR region of 1–3 μm; however, they only capture a part of the region due to a cut-off wavelength of 1.7 μm. This study presents an InAs p-i-n photodetector grown on a GaAs substrate (001) by inserting 730-nm thick InxAl1−xAs graded and AlAs buffer layers between the InAs layer and the GaAs substrate. At room temperature, the fabricated InAs photodetector operated in an infrared range of approximately 1.5–4 μm and its detectivity (D*) was 1.65 × 108 cm · Hz1/2 · W−1 at 3.3 μm. To demonstrate performance, the Sherlock Holmes mapping images were obtained using the photodetector at room temperature.
This content is subject to copyright. Terms and conditions apply.
1
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
InAs on GaAs Photodetectors Using
Thin InAlAs Graded Buers and
Their Application to Exceeding
Short-Wave Infrared Imaging at
300 K
Soo Seok Kang1,2, Dae-Myeong Geum
1, Kisung Kwak1, Ji-Hoon Kang1,
Cheol-Hwee Shim
3, HyeYoung Hyun3, Sang Hyeon Kim
1, Won Jun Choi1,
Suk-Ho Choi2, Min-Chul Park1 & Jin Dong Song
1
Short-wave infrared (SWIR) detectors and emitters have a high potential value in several elds of
applications, including the internet of things (IoT) and advanced driver assistance systems (ADAS), gas
sensing. Indium Gallium Arsenide (InGaAs) photodetectors are widely used in the SWIR region of 1–3
μm; however, they only capture a part of the region due to a cut-o wavelength of 1.7 μm. This study
presents an InAs p-i-n photodetector grown on a GaAs substrate (001) by inserting 730-nm thick
InxAl1xAs graded and AlAs buer layers between the InAs layer and the GaAs substrate. At room
temperature, the fabricated InAs photodetector operated in an infrared range of approximately 1.5–4
μm and its detectivity (D*) was 1.65 × 108 cm · Hz1/2 · W1 at 3.3 μm. To demonstrate performance, the
Sherlock Holmes mapping images were obtained using the photodetector at room temperature.
Indium Gallium Arsenide (InGaAs) photodetectors on Indium Phosphide (InP) substrates are commonly used to
detect the ‘classic’ short-wave infrared (SWIR) range of 1–1.7 μm. Such devices are widely used in spectroscopy,
night vision, gas sensing and telecommunications applications15.
Recently, a ‘capacity crunch’ in this wavelength range was predicted based on the explosive increase in network
trac that is being driven by the development of the internet of things (IoT). Systems such as ber optics, wave-
guides, emitters, and detectors are undergoing extensive research to nd ways to extend performance from 1.7 μm
to longer wavelengths in the SWIR range611. Such extensions could support a variety of applications, including
pedestrian detection for advanced driver assistance systems (ADAS), and the identication of particulate matter
in the air12.
HgCdTe (MCT) detectors are currently in widespread use in the SWIR range, but detectors are subject to
high price and require an integrated cooling system13. Also, the extended InGaAs and InGaAs/GaAsSb type-II
quantum well photodetectors have been intensively studied, but so far cannot detect the full SWIR region1417.
Detectors that are of low-cost and fully cover the SWIR region are required.
Indium arsenide (InAs), one of the III-V materials, has a high electron mobility of 30,000 cm2/V·s and a direct
band gap of 0.417 eV. Due to the narrow band gap, InAs-based detectors can sense SWIR light that is longer than
the cut-o wavelength of a conventional InGaAs photodetector. However, although the physical properties of
InAs constitute a photodetector that is advantageous, manufacturing the detectors is dicult due to the cost of
InAs wafers, which are more expensive than InP wafers.
GaAs substrates are cheaper than InAs and InP substrates; however, the lattice mismatch between GaAs and
InAs is 7.2%, which leads to inevitable defects such as mist dislocations (MDs) and threading dislocations (TDs).
1Center for Opto-Electronic materials and devices, Korea Institute of Science and Technology, Seoul, 136-791, Republic
of Korea. 2Department of Applied Physics and Institute of Natural Sciences, Kyung Hee University, Yongin, 17104,
Republic of Korea. 3Advanced Analysis Center, Korea Institute of Science and Technology, Seoul, 136-791, Republic of
Korea. Correspondence and requests for materials should be addressed to M.-C.P. (email: minchul@kist.re.kr) or J.D.S.
(email: jdsong@kist.re.kr)
Received: 20 May 2019
Accepted: 22 August 2019
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
e dislocations have a negative impact on the physical properties of InAs due to electron-defect scattering. In
order to minimize defects caused by lattice mismatch, it is essential to employ a metamorphic growth technique.
For example, these include the low and high growth temperature (LT-HT) InxGa1xAs step graded buer and
InxAl1xAs step graded buer layer methods1820.
In the LT-HT growth method, the InAs lm is grown at LT and then the InAs lm is grown at HT. e initial
high density InAs nucleation covers the GaAs substrate at LT, and the strain induced by lattice mismatch between
the GaAs and InAs relaxes by generating high density dislocations. Subsequently, the LT InAs is reconstructed
into a continuous lm by heating to HT. en, the reconstructed LT InAs lm acts as a pseudo substrate for the
growth of high-quality InAs lms18. In LT GaAs lms grown on an InP substrate, the dislocations generated by
stress-induced lattice mismatch are immobilized and the dislocations slide towards a symmetric orientation. As
the lm thickness increases, the dislocation density decreases by symmetrically oriented dislocation scattering in
two dimensions21.
In the graded buer layer method, the stress induced by lattice mismatch can be relaxed through stepwise or
linearly constant lattice change, and the dislocation density is decreased due to the hardening of the alloy for an
In composition of 0.5 for InxGa1x(Al1x)As20.
Recently, Loke et al.2224 have reported an InAs photodetector grown on a GaAs substrate using the InxAl1xAs
graded buer layer and LT-GaAs. e composition x was continuously changed from 0 to 1. e InAs layer was
the surface of the graded buer layer, and was not an insulating layer for the InAs lm. is conductive layer pre-
vented the precise electronic characterization of the InAs layer. Moreover, the motion of dislocations, which leads
to the degradation of the devices due to carrier-defect scattering, is not directly shown in this report.
In our previous work, a high-quality InSb lm was grown on a GaAs substrate using an InxAl1xSb continuous
graded buer layer25. Unlike other reports, the growth temperature of the continuous graded buer layer grad-
ually decreased because the growth temperature of InSb is lower than that of AlSb. Proper control of the growth
temperature can minimize slippage and generation of dislocations by increasing yield strength.
In the aforementioned two references, the thickness of the continuous buer layer was above 1.4 μm. e
dislocation density is well known to be inversely proportional to lm thickness. However, a thick buer layer is
not desirable for real applications.
In this work, a high-quality InAs lm was obtained using an InxAl1xAs graded buer layer, with x = 0 to
x = 0.87 for the insulating surface of the graded buer layer. e composition x was gradually changed by varying
the growth rate of In and Al, while gradually decreasing the growth temperature. e In0.87Al0.13As layer closest
to the InAs lm showed sucient insulating property to allow precise electronic measurement of the InAs lms,
and has proper Al composition to be used as sacricial layer for future integration. In our study, the eect of the
InxAl1xAs graded buer layer for fabricating high-quality InAs lm is better than that of LT-HT buer layers.
To fabricate a SWIR imaging operating at room temperature, the p-i-n InAs was grown on a semi-insulating (SI)
GaAs substrate by inserting the InxAl1xAs graded buer layer.
Results and Discussion
ree samples of InAs were grown on SI-GaAs (001) using In0.87Al0.13As - In0.87Al0.13As (LH1), InAs - In0.87Al0.13As
(LH2), and InxAl1xAs graded (G1) buer layers in a Riber compact 21E solid source molecular beam epitaxy
(MBE) system, as shown in Fig.1a.
In the LH1 and LH2 samples, the GaAs substrate was degassed at 400 °C and then the removal of the native
oxide from the substrate was carried out in an As2 atmosphere supported by a valved-cracker As cell at the sub-
strate temperature (Ts) of 620 °C for 10 min. Subsequently, a 100-nm thick GaAs buer layer having a Ga growth
rate of 1.82 Å/s was deposited on a at surface at 580 °C. An InAs layer of 300 nm thickness was grown in a GaAs
buer layer at the LT of 270 °C and then an In0.87Al0.13As layer of 300 nm thickness was cultured at an HT of 470 °C
in the LH1. Subsequently, an InAs lm having a thickness of 500 nm was grown at 470 °C. e growth procedure
of LH2 is the same as that of LH1, but with a 300-nm thick In0.87Al0.13As layer, instead of an InAs lm grown at
LT. e In growth rate of In0.87Al0.13As and InAs was 2.62 Å/s, while the Al growth rate of the LT-HT samples was
0.4 Å/s.
During the growth procedure of the G1 samples, the degassed, deoxy, and GaAs buer layer deposition pro-
cesses were similar to those mentioned above. en, a 100 nm thick AlAs layer was deposited with an Al growth
rate of 2.7 Å/s at a Ts of 580 °C. At that point, an InxAl1xAs graded buer layer was grown while gradually chang-
ing the growth rate of In from 0 Å/s to 2.62 Å/s, and the Al from 2.7 Å/s to 0.4 Å/s, while the Ts was gradually
decreased from 530 to 470 °C. Finally, a 500-nm InAs lm was grown at a Ts of 470 °C.
e structure of the In0.87Al0.13As terminated (GT) sample is shown in Fig.1a, and the sample demonstrated
an insulating property.
Figure1b shows the electron mobilities and surface roughness of four samples. e electron mobilities of LH1,
LH2, G1, and GT were 4,375, 4,749, 9,275 and 617 cm2/V·s, respectively. Apart from the GT sample, the other
samples showed the electron mobilities of a 500-nm thick InAs lm, and the G1 sample had the highest electron
mobility. is implies that the graded buer layer can overcome the lattice mismatch between GaAs and InAs,
and that it is a more ecient approach than the LT-HT growth method for obtaining a high quality InAs lm on
a GaAs substrate.
e surface roughness of LH1, LH2, G1, and GT was 1.1, 1.5, 2.1, and 7.2 nm, respectively, and obtained
images of 10 μm × 10 μm, as shown in Fig.1c. Although the surface roughness of G1 was higher than that of
LH1 and LH2, the electron mobility of G1 was also higher than that of LH1 and LH2. is means that surface
scattering does not have a major inuence in the degradation of carrier transport, taking into account the electron
mobilities and surface roughness.
e degradation of electron mobility ascribes to electron-defect scattering, and several hillocks and pits in
LH1 and LH2 are slightly higher than that in G1, as shown in Fig.1c. e high density of hillocks and pits has
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
a negative impact on electron transport. e surface morphology of GT is the roughest among the samples,
and the pyramidal surface morphology implies that the graded buer layer could grow at low growth tempera-
ture. Interestingly, the high surface roughness of the graded buer layer reduced from 7.2 nm to 2.1 nm aer the
500 nm thick InAs grew in the GT. e electron mobility of the GT was 617 cm2/V·s, and its resistivity of 18 /
cm was also higher than the G1 resistivity of 0.015 /cm (data not shown). erefore, the electron mobility of the
500-nm thick InAs lm was precisely achieved due to the insulating property of the GT layer.
e 2θ/ω spectra of samples are shown in Fig.1d. e In0.87Al0.13As 2θ/ω peaks in LH1 and LH2 were at
61.71° and 61.94°, respectively, while the GaAs (004) peak was 66.05°. e 2θ/ω angle dierence of In0.87Al0.13As
between LH1 and LH2 is attributed to the In and Al growth rates, which was measured at high temperature. e
growth rate varied at low temperature. In particular, the In growth rate rapidly increased due to a reduction in the
re-evaporation of In at low temperature.
For G1, the peak of the InxAl1xAs graded buer layer was widely obtained in the range of 65.5° to 61.85°. e
InAs (004) peak and its full width at half maximum (FWHM) are shown in Fig.1e. e InAs (004) peaks of LH1,
LH2, and G1 were positioned at 61.08°, 61.04°, and 61.05° and the lattice constant from the peaks was 6.064 Å,
6.067 Å, and 6.066 Å, respectively. e lattice constants are slightly larger than those of InAs, at 6.058 Å, and the
epitaxial InAs lms are almost relaxed. e threading dislocation density of the InAs lm in the G1 sample is
9.4 × 107/cm2, which was calculated by Ayer’s model in XRD spectra.
e 500 nm thick InAs lm in G1 was lattice mismatched to the GaAs by 7.32% out of plane. is means that the
lattice of InAs lms was stretched out of the plane. e FWHM of G1 was wider than that of LH1 and LH2, while
the G1 sample showed the highest electron mobility among these samples. e electron-defect scattering of the G1
sample was minimized due to the relatively low defect density in these samples. erefore, the stress induced by the
Figure 1. Growth schemes of InAs lms on a GaAs substrate and their structural and electrical properties. (a)
Schemes of sample growth using In0.87Al0.13As - In0.87Al0.13As (LH1), InAs - In0.87Al0.13As (LH2), InxAl1xAs
graded (G1), and In0.87Al0.13As terminated (GT) buer layers on GaAs substrates. (b) Electron mobility and
surface roughness of the samples. (c) Surface morphology, obtained atomic force microscopy (AFM), of the
samples. e scale bar is 5 μm. (d) XRD 2θ/ω spectra of the samples. e dashed line indicates the InAs (400)
peak position. (e) InAs (400) 2θ/ω peak and its full width at half maximum (FWHM).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
lattice mismatch between InAs and GaAs relaxes by the deformation of the atomic structure of InAs in the G1 sam-
ple, while the stress relaxes by sliding and the generation of dislocations in the LH1 and LH2 samples.
e InAs p-i-n photodetector sample grew on an InxAl1xAs graded buer layer in the G1 scheme. e p- and
n-type InAs lm grew with Be and Si at a doping concentration of 3 × 1018/cm3 and 5 × 1017/cm3, respectively. e
structural scheme and the TEM dark-eld image of the p-i-n InAs lm are shown in Fig.2a,b, respectively. e
AlAs and InxAl1xAs layers are not clearly distinguishable, but the thickness of the layers in total is approximately
730 nm. Also, a high density of defects is shown in the layers.
Meanwhile, the thickness of the InAs p-i-n layer was 2370 nm, and some edge dislocations appeared as defects in
the TEM image. erefore, the graded buer layer suppressed the nucleation and sliding of the dislocations, inducing a
lattice mismatch between the InAs lm and the GaAs substrate, and the dislocations connes to the graded buer layer.
e brightness in a HAADF image indicates the number of electrons scattered inelastically by the atoms in a
sample, and the heavy atoms appear to be brighter than the light atoms. Figure2c exhibits a high angle annular
dark eld (HAADF) image of the InAs p-i-n sample with an AlAs layer of ~110 nm, and distinctly shows the
InAlAs layer of ~620 nm in thickness due to the dierence in brightness. Furthermore, the brightness increases
from the surface of the AlAs layer to that of the InxAl1xAs graded buer layer, which implies that the Al and In
compositions have gradually changed in the buer layer. e white circles dissipate In rich droplets during the
focused ion beam (FIB) milling process in the InAs p-i-n region. e interface between the InxAl1xAs graded
buer and the InAs p-i-n layer is rugged and corresponds to the pyramidal shaped surface of the GT AFM image,
as shown in Fig.1c.
e limited In diusion results in a dierence in growth rates along the (001) and (111) faces due to the
atomic density in these directions. e low growth temperature limits dislocation motion and increases yield
strength26,27. us, structural deformation is more benecial than the sliding and nucleation of the dislocations to
mitigate stress caused by the lattice mismatch and induced by the low growth temperature. Similarly, the disloca-
tions reect in the interface between the InAs and InAlAs graded buer layer.
e TEM image of the InAs/InxAl1xAs graded buer/AlAs/GaAs, obtained by two beam conditions, is shown in
Fig.2d, and the black circles in the InAs are the In rich droplets. e mist and threading dislocations appear in the
lower region of the InxAl1xAs graded buer layer, and the dislocations generated in the lower region have slipped.
In this case, the dislocation density decreases with increasing lm thickness. Meanwhile, the dislocation density in
the upper region of the InxAl1xAs graded buer layer decreases as the InAs layer closes through threading disloca-
tion blocking by mist dislocation segment and fusion process. In addition, the dislocations in the upper region of
the InxAl1xAs graded buer layer almost do not slip directly to the InAs layer. In other words, the dislocations are
conned in the InxAl1xAs graded buer layer. e localization of the dislocations is attributed to the abrupt inter-
face28,29. e connement of dislocations is also shown in Fig.2b, where point A is explained by Fig.2f.
Figure 2. Structural diagram and micro structure analysis of p-i-n inas on a GaAs substrate. (a) Scheme of the
p-i-n InAs on a GaAs substrate. (b) Transmission electron microscopy (TEM) dark-led image of p-i-n InAs
on a GaAs substrate. (c) High angle annular dark eld (HAADF) image of p-i-n InAs on a GaAs substrate. (d)
High-resolution TEM image of the InAs/InAlAs graded buer/AlAs/GaAs substrate. e A point indicates
the change in the rate of increase in the in-plane strain, as shown in the strain line prole (f). e image shows
threading dislocations blocked by mist dislocation segments, and threading dislocation fusions in white
circles. (e) Energy-dispersive X-ray spectroscopy (EDS) elemental line prole of p-i-n InAs on a GaAs substrate.
(f) In-plane and out-of-plane strain line prole of p-i-n InAs on a GaAs substrate. e A point indicates the
point in the TEM image (d).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
e normalized EDS line prole is illustrated in Fig.2e. e amount of Al and In gradually decreases and
increases, respectively, along the growth direction, as expected with an InxAl1xAs graded buer layer. e change
in the In and Al composition in the graded buer layer is consistent with the increased brightness in the HADDF
image shown in Fig.2c. As noted, this brightness relates to the atomic mass.
e strain line prole of the InAs p-i-n sample is shown in Fig.2f and is relatively obtained to the lattice
constant of GaAs substrate by the Topspin experiment. e inset of Fig.2f shows the enlarged strain prole of
the InxAl1xAs layer in the sample. Here, sxy and szz indicate the in-plane (110) and out-of-plane (001) strain,
respectively. e szz in the AlAs layer increases slightly, while sxy has almost no change. is means that the stress
induced by the lattice mismatch between the GaAs and AlAs is alleviated by the deformation of the AlAs cubic
structure. e sxy and szz gradually increase in the InxAl1xAs graded buer layer, and the increased rate of strain
is conformable, from the AlAs to point A.
e similar increase in the rates of sxy and szz in the lower region of point A attributes to stress relaxation
caused by the generation and sliding of dislocations. On the other hand, the szz rate of increase is faster than that
of sxy in the upper region of point A. e dierence in rates between sxy and szz is attributed to the deformation of
the InxAl1xAs cubic structure. erefore, the stress in the upper region of point A is dominantly relieved by the
deformation of the cubic structure rather than the generation and sliding of dislocations. In other words, for stress
relaxation, the generation and sliding of the dislocations is advantageous in the lower part of point A, while the
deformation of the cubic structure is advantageous in the upper part of point A.
At the interface between the graded buer layer and the InAs layer, szz and sxy are 7.45% and 7%, respec-
tively. e szz is in good agreement with the XRD result, which is approximately 7.32%. As the thickness of InAs
increases, the dierence between the sxy and szz strain gradually decreases until the szz and sxy are about 7.45%
and 7.35%, respectively, at the surface of the InAs layer. Although the strain of InAs against GaAs is larger than
the original strain of 7.25%, the gradual equalization of the strains implies that the atomic structure of InAs has
recovered from the deformed cubic to the cubic form.
Figure3 shows the electrical and opto-electrical properties of the fabricated detector. Figure3a exhibits
the current density versus bias (J-V) curve of the detector measured at room temperature, and shows the
rectication characteristics of the p-i-n junction. e dark current density of the InAs photodetector, meas-
ured at room temperature under a bias of 0.5 V, is 4.6 A/cm2 as shown in Fig.3b. It is slightly lower than
those of other InAs photodetectors (PDs) on a GaAs substrate. e dark current densities of InAs PDs, which
were grown using a LT-InAs and an InAlAs graded buer layer, are 9.4 A/cm2 and 7 A/cm2, respectively22.
Furthermore, it is worth noting that these values are approximately two orders of magnitude higher than that
of 0.064 A/cm2 from a homo-epitaxial InAs photodetector due to the dependence of the dark current density
on the dislocation density30.
Figure3c shows the normalized photoresponse of the detector as a temperature variation, and the cut-o
wavelength is red-shied from about 3.1 μm at 77 K to 4 μm at 300 K. e changed cut-o wavelengths relates to
the band gap of InAs at various temperatures. e peak responsivity is shown in Fig.3d and its value varies from
0.6 A/W at 80 K to 0.126 A/W at 300 K. e detectivity (D*) of the photo-detector is given by
=
DR
RA
kT4(1)
p0
1/2
where k, T, and A are Boltzmann’s constant, temperature, and the window of the device, respectively.
e D* of the InAs photodetector at room temperature is estimated by Eq. (1) to be 1.65 × 108 cm · Hz1/2 · W1
at 3.3 μm.
In order to verify the imaging capability of the InAs photodetector, 2D image scanning was performed by
means of a bi-axial mechanical scanner. Figure3e depicts a schematic diagram of a single-pixel imaging system
in a transmissive conguration used for 2D imaging in the short wave infrared (SWIR) range. Since the InAs
photodetector was used as a single-pixel sensor, a bi-axial mechanical scanner was used to obtain 2D images.
Also, two linear stages (Physik Instrument, Q-545) were used to scan the x and y axes, respectively. e 2D
photo-detection mapping image was obtained using a ltered IR-light source with a band-pass lter, and an open-
work thin Sherlock Holmes bookmark was used as a scanning object. e position of the scanned object was in
the plane where the image of the pinhole spot was formed by the use of lenses. erefore, the output images can
be obtained by scanning the spot through the scanned object. e scanned object was attached to the x-y planar
moving stage, and the object was scanned and captured by the single-pixel InAs photodetector. A source meas-
urement unit (Keysight Technologies, B2901A) was used to measure photocurrents. Each pixel value corresponds
to the current measurement obtained at the corresponding scanner position.
e 2D mapping images of the photo-detection are shown in Fig.3f. A band-pass lter with an FWHM of
0.5 ± 0.1 μm lters the light source. e 2D mapping image with a 2250 nm lter is brighter than that of a 2500 nm
lter, but a stronger photoresponse is attributed to the spectral dependence of the light-source power. Despite the
dierences in photoresponse, the Sherlock Holmes image is clearly displayed, which means that the InAs p-i-n
photodetector, grown using the InxAl1xAs graded buer layer, can be used as an image sensor for the MIR region.
Conclusion
To obtain a high quality InAs lm, we investigated an InAs lm grown on a GaAs substrate using a LT-HT and
graded buer layer. e InAs lm with the graded buer layer exhibited the highest electron mobility among the
lms and its value is 9,275 cm2/V·s. e InAs p-i-n photodetector on a GaAs substrate grew using an InxAl1xAs
graded buer layer (x = 0 0.87). e cut-o wavelength of the detector is approximately 4 μm and the respon-
sivity is 0.126 A/W at room temperature.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
Figure 3. Performance of the InAs photodetector. (a) Current density-voltage (J-V) curve of the fabricated
InAs photodetector, measured at room temperature. (b) Dark current density values reported by others, at a
bias of – 0.5 V at room temperature. (c) Photoresponse of the InAs photodetector at various temperatures. e
cut-o wavelength is simultaneously red-shied by increasing temperature. (d) Peak responsivity of the InAs
photodetector at various temperatures, measured by a blackbody temperature of 700 °C and chopper frequency
of 500 Hz. e detectivity (D*) is obtained using Eq. (1). (e) Schematic illustration of the single-pixel imaging
system in a transmissive conguration used for 2D imaging in the short-wave infrared range. (f) 2D photo-
detection images at the peak wavelengths of 2.25 μm and 2.5 μm, ltered by band-pass lter.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
In future work, we will focus on the research of the integration of InAs photodetectors with wafer bond-
ing and epitaxial li-o method, because the InAlAs/AlAs graded buer was designed for selective etching31.
Additionally, the InAs photodetector will be optimized.
Methods
Electrical properties such as resistivity, carrier mobility, and carrier concentration were measured by an Ecopia
HMS-3000 Hall Measurement System with an input current of 0.1 mA at room temperature.
The InAs surface and morphology were measured using a Park systems XE-100 with an NSC-36 tip in
non-contact mode.
e lattice spacing on the out-of-plane samples was measured using a Rigaku ATX-G XRD.
e microstructure was studied using a TECNAI F20 G2 SuperTwin TEM (FEI, Hillsboro, OR) with the
sample mounted in a double-tilt holder (Gatan, Pleasanton, CA). TEM images were obtained by the Fischhione
Model 3000 ADF detector and UltraScan 1000 (2k × 2k) CCD camera (Gatan Inc.). Energy dispersive spectros-
copy (EDS) and TopSpin analysis, which are strain analysis tools, were performed using a Talos F200X TEM with
Bruker Super EDS system and NanoMEGAS DigiSTAR precession electron diraction (PED) unit with Appve
soware in TENCAI F20 G2 SuperTwin TEM, respectively.
e current density of the 500 × 500 μm detector was measured at 300 K by a Keithley 4200 integrated in a
probe station. e photocurrent spectra were obtained using a customized system with a Bruker VERTEX 80 v
Fourier transform infrared spectrometer, a global MIR source, a Janis cryo-chamber, and a Keithley 428 low-noise
current amplier. e peak responsivity was measured using a fully-integrated smart home system with a 700 °C
blackbody source, vacuum chamber, and SR830 lock-in amplier for a 77 K low temperature experiment.
For 2D imaging, the optical wavelength range from an IR-source (SLS202L, Thorlab) was filtered by a
band-pass lter (FB2250-500 or FB2500-500, orlab), and then the optical path was passed to the single InAs
p-i-n photodetector or masked by a Sherlock Holmes bookmark supported by an XY automated stage. e exper-
iment was carried out at room temperature.
References
1. Bachmann, . J. & Shay, J. L. An InGaAs detector for the 1.0–1.7-µm wavelength range. Appl. Phys. Lett. 32, 446–448 (1978).
2. MacDougal, M. et al. Low dar current InGaAs detector arrays for night vision and astronomy. Proc. SPIE. 7298,
72983F-1–72983F-10 (2009).
3. Nameata, N., Sasamori, S. & Inoue, S. 800 MHz single-photon detection at 1550-nm using an InGaAs/InP avalanche photodiode
operated with a sine wave gating. Opt. Express 14, 10043–10049 (2006).
4. Nameata, N., Adachi, S. & Inoue, S. 1.5 GHz single-photon detection at telecommunication wavelengths using sinusoidally gated
InGaAs/InP avalanche photodiode. Opt. Ex press 17, 6275–6282 (2009).
5. van Well, B. et al. An open-path, hand-held laser system for the detection of methane gas. J. Opt. A-Pure. Appl. Op. 7, S420–S424
(2005).
6. Zhenwei, X. et al. Ultra-broadband on-chip twisted light emitter for optical communications. Light-Sci. Appl. 7, 18001 (2018).
7. Dianov, E. M. Bismuth-doped optical bers: a challenging active medium for near-I lasers and optical ampliers. Light-Sci. Appl.
1, e12 (2012).
8. aphael, F. et al. Shaping the light amplied in a multimode ber. Light-Sci. Appl. 6, e16208 (2017).
9. Hou, C.-C. et al. Near-infrared and mid-infrared semiconductor broadband light emitters. Light-Sci. Appl. 7, 17170 (2018).
10. Singh., N. Octave-spanning coherent supercontinuum generation in silicon on insulator from 1.06 µm to 2.4 µm. Light-Sci. Appl. 7,
17131 (2018).
11. Acert, J. J. et al. High-speed detection at two micrometres with monolithic silicon photodiodes. Nat. Photonics 9, 393–396 (2015).
12. Singh, A. et al . Indoor air pollution and its association with poor lung function, microalbuminuria and variations in blood pressure
among itchen worers in India: a cross-sectional study. Environ. Health 16, 33 (2017).
13. Simingalam, S. et al. Development and fabrication of extended short wavelength infrared HgCdTe sensors grown on CdTe/Si
substrates by molecular beam epitaxy. Solid State Electron. 101, 90–94 (2014).
14. Hoogeveen, . W. M., onald, J. V. D. A. & Goede, P. H. Extended wavelength InGaAs infrared (1.0-2.4 µm) detector arrays on
SCIAMACHY for space-based spectrometry of the Earth atmosphere. Infrared Phys. Techn. 42, 1–16 (2001).
15. Yoon, H. W., Dopiss, M. C. & Eppeldauer, G. P. Performance comparisons of InGaAs, extended InGaAs, and short-wave HgCdTe
detectors between 1 µm and 2.5 µm. Proc. SPIE. 6297, 629703-1-629703-10 (2006).
16. Chen, B., Jiang, W., Yuan, J., Holmes, A. L. Jr. & Onat, B. M. SWI/MWI InP-based p-i-n photodiodes with InGaAs/GaAsSb type-
II quantum wells. IEEE J. Quantum Electron. 47, 1244–1250 (2011).
17. Uliel, Y. et al. InGaAs/GaAsSb type-II superlattice based photodiodes for short wave infrared detection. Infrared Phys. Technol. 84,
63–71 (2017).
18. Yuan, H., Chua, S. J., Miao, Z., Dong, J. & Wang, Y. Growth and structural properties of thic InAs lms on GaAs with low-pressure
metalorganic vapor phase epitaxy. J. Cryst. Growth 273, 63–67 (2004).
19. Chang, C.-A., Serrano, C. M., Chang, L. L. & Esai, L. Studies by cross-sectional transmission electron micro scope of InAs grown
by molecular beam epitaxy on GaAs substrate. Appl. Phys. Lett. 37, 538–540 (1980).
20. Jeong, Y., Choi, H. & Suzui, T. Invalidity of graded buers for InAs grown on GaAs (001)-A comparison between direct and
graded-buer growth. J. Cryst. Growth 302, 235–239 (2007).
21. Mathis, S. ., Wu, X. H., omanov, A. E. & Spec, J. S. reading dislocation reduction mechanisms in low-temperature-grown
GaAs. J. Appl. Phys. 86, 4836–4842 (1999).
22. Loe, W. ., Tan, . H., Li, D., Wicasono, S. & Yoon, S. F. Mid-infrared InAs photodetector grown on GaAs substrate through
cation exchange. IEEE Photon. Technol. Lett. 29, 458–461 (2017).
23. Loe, W. ., Tan, . H., Li, D., Wicasono, S. & Yoon, S. F. oom temperature 3.5-m mid-infrared InAs photovoltaic detector on a
Si substrate. IEEE Photon. Technol. Lett. 28, 1653–1656 (2016).
24. Loe, W. ., Tan, . H., Wicasono, S. & Yoon, S. F. Epitaxy growth and characterization of InAs p-i-n photodetector through ion
exchange for mid-infrared detection on Si substrates. MS Commun. 8, 1085 (2018).
25. Shin, S. H., Song, J. D., Lim, J. Y., oo, H. C. & im, T. G. Structural and electrical properties of high-quality 0.41 µm-thic InSb
lms grown on GaAs (001) substrate with InxAl1xSb continuously graded buer. Mater. es. Bull. 47, 2927–2930 (2012).
26. I. Yonenaga, Mater. Trans. 46, 1979 (2005).
27. Wang, H. et al. Eects of growth temperature on highly mismatched InAs grown on GaAs substrates by MBE. J. Cryst. Growth 186,
38–42 (1998).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
SCIENTIFIC REPORTS | (2019) 9:12875 | https://doi.org/10.1038/s41598-019-49300-z
www.nature.com/scientificreports
www.nature.com/scientificreports/
28. Dimroth, F. et al. Metamorphic GaInP/GaInAs/Ge triple-junction solar cells with> 41% eciency. 34th IEEE Phot. Spec. Conf.
001038–001042, (2009).
29. Jandl, A., Bulsara, M. T. & Fitzgerald, E. A. Materials properties and dislocation dynamics in InAsP compositionally graded buers
on InP substrates. J. Appl. Phys. 115, 153503 (2014).
30. Dobbelaere, W. et al. InAs p-n diodes grown on GaAs and GaAs-coated Si by molecular beam epitaxy. Appl. Phys. Lett. 60, 868–870
(1992).
31. Geum., D.-M. et al. Ultra-high-throughput production of III-V/Si wafer for electronic and photonic applications. Sci. ep. 6, 20610
(2016).
Acknowledgements
is work was mainly supported by KIST Institutional Program (2E29390).is work was partially supported
byNRFgrant funded by the Korea government(MSIP) (No. 2019M3F3A1A02072069).
Author Contributions
Jin Dong Song managed the research and supervised the experiment. Soo Seok Kang grew the samples and
analyzed the experiments. Suk-Ho Choi proofread the manuscript. Dae-Myeong Geum, Sang Hyeon Kim and
Won Jun Choi fabricated and analyzed performance of the InAs photodetector. Cheol-Hwee Shim and HyeYoung
Hyun carried out TEM measurements. Kisung Kwak and Min-Chul Park carried out and analyzed the 2D
imaging experiment. All authors reviewed the manuscript. Ji-Hoon Kang provided exact expression and modied
gure 3e of image scanning.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-019-49300-z.
Competing Interests: e authors declare no competing interests.
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) 2019
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
... Before the growth of the InAs PD active layer, we first Comparison of threading dislocation densities in InAs/GaAs materials reported by other groups. 21,23,24,27,28 desorbed the GaAs substrate at 540°C for 10 min under As 2 overpressure to remove native oxide. Following oxide desorption, the 50 nm GaAs and AlAs sacrificial layers were grown at 580°C. ...
... We compared our low TDD values with previously reported studies (Figure 2g), showing that Sample B is 5 × lower than the previously lowest TDD. 21,23,24,27,28 This record-low TDD value was achieved using both the well-optimized In x Al 1−x As graded buffer and DFL. Figure 3a shows the fabrication procedure of our flexible InAs PD arrays. Standard photolithography was used to define 20 × 10 mesa arrays. ...
... Table 1 compares our result to the various InAs PDs reported by other groups. 28,31,37,38 This work not only shows the first flexible InAs mid-wavelength infrared PD arrays but also reveals the highest specific detectivity among the heteroepitaxially grown InAs PDs. For future work, further optimization of In 0.95 Al 0..5 As DFLs and the structure for efficient cavity effect will be done to further improve the PD performance. ...
Article
Full-text available
Current infrared thermal image sensors are mainly based on planar firm substrates, but the rigid form factor appears to restrain the versatility of their applications. For wearable health monitoring and implanted biomedical sensing, transfer of active device layers onto a flexible substrate is required while controlling the high-quality crystalline interface. Here, we demonstrate high-detectivity flexible InAs thin-film mid-infrared photodetector arrays through high-yield wafer bonding and a heteroepitaxial lift-off process. An abruptly graded InxAl1–xAs (0.5 < x < 1) buffer was found to drastically improve the lift-off interface morphology and reduce the threading dislocation density twice, compared to the conventional linear grading method. Also, our flexible InAs photodetectors showed excellent optical performance with high mechanical robustness, a peak room-temperature specific detectivity of 1.21 × 10⁹ cm-Hz1/2/W at 3.4 μm, and excellent device reliability. This flexible InAs photodetector enabled by the heteroepitaxial lift-off method shows promise for next-generation thermal image sensors.
... 18−20 After that, the lattice mismatch is gradually relaxed through the introduction of dislocations. Therefore, this metamorphic growth normally requires complicated growth sequences or very thick layer growth (several μm), such as graded-buffer growth, 13,16,17 to realize high-quality 2D layers with flat surfaces. ...
... Such an interface structure is characteristic of the InAs growth on GaAs (111)A, 21−24 and it is difficult to realize the structure on GaAs (100). 17 Figure 5 shows the responsivity spectra of the device at 77 K with various applied voltages. The photo-response was measured by the lock-in method with a chopping frequency of ∼3 kHz. ...
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.
... Indium gallium arsenide (InGaAs) and indium aluminum arsenide (InAlAs) are ternary group III-V semiconductor compounds, which have wide ranges of optoelectronic applications. The excellent lattice matching of InGaAs and InAlAs with InP (In 0.53-Ga 0. 47 As and In 0.52 Al 0.48 As) are commonly used to make infrared photodetectors, which are used in telecommunications, night vision, and pedestrian detection in advance driver assistance systems [1][2][3][4][5]. Due to the high carrier mobility and electrical conductivity of these semiconductors, these photodetectors have very short response time (rt \ 1 ns), high quantum efficiency (g Q ext [ 50%), and low dark current (i D \ 10 -8 A) [2,6]. ...
Article
Full-text available
The complex dielectric function (ε = ε1 + iε2) spectra of epitaxial In0.53Ga0.47As and In0.52Al0.48As in multilayer stacks were extracted by spectroscopic ellipsometry. Unlike previous studies over more limited spectral ranges, these analyses span from the THz to ultraviolet and provide a non-contacting method of determining optical and electronic transport properties from ε of different layers. The direct band gap of these In0.53Ga0.47As and In0.52Al0.48As films were determined to be 0.73 and 1.44 eV, respectively. Above band gap critical point transition energies were measured to be 1.22, 2.53, and 2.87 eV for In0.53Ga0.47As and 1.74 and 2.95 eV for In0.52Al0.48As, respectively. Biaxial tensile stress in epitaxial In0.53Ga0.47As and In0.52Al0.48As films were measured to be 0.15 and 0.11 GPa, respectively. The amplitude of phonon modes in In0.53Ga0.47As and In0.52Al0.48As at 252 and 351 cm⁻¹ were determined to be 105 and 99, respectively. Resistivity (ρ), scattering time (τ), carrier mobility (μ), and carrier concentration (N) were determined using the Drude model, yielding ρ = 0.00412 Ωcm, τ = 238 fs, μ = 9975 cm²/Vs, and N = 1.52 × 10¹⁷ cm⁻³ for In0.53Ga0.47As and ρ = 0.026 Ωcm, τ = 189 fs, μ = 4435 cm²/Vs, and N = 5.3 × 10¹⁶ cm⁻³ for In0.52Al0.48As. The transport properties measured from ellipsometry are in close agreement with Hall effect measurement and reported results in literature. Optoelectronic parameters from these analyses can serve as input to model In0.53Ga0.47As and In0.52Al0.48As based optoelectronic devices to optimize their performance and to understand device physics.
... Therefore, efforts have been channeled towards improving the efficiency and the performance of the photodetector. In previous studies, the improvement to enhance the efficiency was mainly focused on the device's active area [2,3]. However, a relatively complicated process is required for the fabrication of the device. ...
Article
Full-text available
In this project, the surface structure of III-V semiconductor, GaAs, was altered to enhance the optical and electronic properties of the semiconductor. This project involved the designing and fabrication of non-porous and porous GaAs structures using SILVACO TCAD tools. The porous GaAs with different pore depth were designed and simulated to investigate the effect of pore depth on the optical and electrical properties of GaAs semiconductor. The pore depth of porous GaAs structure was varied with 2, 4, 6 and 8 μm. The porous GaAs structures were then tested for the metal-semiconductor-metal (MSM) photodetector device application. The non-porous and porous GaAs MSM photodetectors were compared systematically through current-voltage (I-V) characteristics, current gain, and spectral response. The result showed that the porous GaAs MSM photodetector has better performance in terms of electrical and optical properties than the non-porous photodetector. Amongst the MSM GaAs photodetectors, the porous GaAs photodetector with pore depth of 6 μm obtained the highest current gain value of 3.22. While for optical properties, the spectral response showed the current intensity of 11.370 µA which was recorded at the peak wavelength of 880 nm. Therefore, porous GaAs showed good potential and can be used for optoelectronic device applications such as MSM photodetector.
... InAs as a special case of InAs1-xSbx alloy seems to be particularly important. A better understanding of its basic physical properties is currently the subject of intensified research [8,9]. In this context, a fundamental role is played by a deep understanding of the InAs structure of energy bands, especially for valence bands. ...
Article
Full-text available
Careful selection of the physical model of the material for a specific doping and selected operating temperatures is a non-trivial task. In numerical simulations that optimize practical devices such as detectors or lasers architecture, this challenge can be very difficult. However, even for such a well-known material as a 5 μm thick layer of indium arsenide on a semiinsulating gallium arsenide substrate, choosing a realistic set of band structure parameters for valence bands is remarkable. Here, the authors test the applicability range of various models of the valence band geometry, using a series of InAs samples with varying levels of p-type doping. Carefully prepared and pretested the van der Pauw geometry samples have been used for magneto-transport data acquisition in the 20–300 K temperature range and magnetic fields up to ±15 T, combined with a mobility spectra analysis. It was shown that in a degenerate statistic regime, temperature trends of mobility for heavy- and light-holes are uncorrelated. It has also been shown that parameters of the valence band effective masses with warping effect inclusion should be used for selected acceptor dopant levels and range of temperatures.
Article
Full-text available
Shortwave infrared (SWIR) imaging devices have gained increasing importance in various fields, including security monitoring, biomedical imaging, advanced driver assistance, and semiconductor industry. However, complex technological processes of epitaxial growth and flip‐chip bonding hinder the development of high‐performance and low‐cost SWIR imaging devices. Herein, a novel organic–inorganic hybrid optical up‐conversion (OUC) imaging device by stacking a phosphorescent organic light‐emitting diode (OLED) with a NiSix/Si Schottky barrier diode (SBD) is developed. The pyramidal microstructures of silicon are utilized to greatly enhance light absorption of the NiSix/Si SBD, enabling the device to respond to a broadband SWIR light of 1064, 1310, and 1550 nm that are beyond the bandgap limit of silicon. Significantly, the device demonstrates excellent up‐conversion imaging behaviors at SWIR light with an ultra‐fast refresh rate of over 3000 Hz and a high‐resolution imaging capability of 508 ppi. This work paves the way toward the fabrication of high‐performance, low‐cost silicon‐based OUC devices for SWIR imaging applications.
Article
Metal halide perovskites (MHPs) have received massive research attention because of their unique inherent photophysical properties that drive them for potential application in the field of photoelectric detection. Among them,...
Article
Full-text available
A high‐speed and broadband 5 × 5 photodetector array based on MoS2/In0.53Ga0.47As heterojunction is successfully demonstrated to take full advantage of the type‐II band‐aligned multilayer MoS2/In0.53Ga0.47As. The fabricated devices exhibit good uniformity in the Raman spectrum and clear rectifying characteristics. The fabricated MoS2/In0.53Ga0.47As photodetectors show good optical performances at a broad wavelength range showing high responsivities corresponding to the detectivity of ≈10¹⁰ Jones at −3 V for the incident broadband light from 400 to 1550 nm. A very fast photoresponse is also obtained with a small rise/fall time in the order of microseconds both for visible (638 nm) and shortwave infrared (1310 nm). Finally, the image scanning properties of MoS2/In0.53Ga0.47As devices are demonstrated for visible and infrared light, indicating that the suggested device is one of the promising options for future broadband imager, which can be integrated on the focal plane arrays (FPAs).
Article
Full-text available
We demonstrate a low threading dislocation density (TDD) and smooth surface InAs layer epitaxially grown on Si by suppressing phase separation of InxAl1−xAs (x = 0 to 1) graded buffer and by inserting a tensile-strained In0.95Al0.05As dislocation filter layer. While keeping the total III–V layer below 2.7 μm to avoid thermal cracks, we have achieved a sixfold reduction of TDD in InAs on Si compared to the unoptimized structure. We found a strong correlation between the metamorphic InAs surface roughness and TDD as a function of InxAl1−xAs buffer thickness. An optimal thickness of 175 nm was obtained where both phase separation and 3D islanding growth were suppressed. Moreover, a tensile-strained In0.95Al0.05As dislocation filter layer and high growth temperature of the InAs cap layer further assisted the dislocation reduction process, which led to a TDD to 1.37 × 10⁸ cm⁻². Finally, an InAs p-i-n photodetector grown on the optimized InAs/Si template confirmed its high quality by showing an improved responsivity from 0.16 to 0.32 A/W at a 2 μm wavelength.
Article
We report a comparative study of metamorphic InAs p-i-n photodetectors epitaxially grown on GaAs and Si by molecular beam epitaxy. Linearly graded InAlAs buffers were employed to bridge the high lattice mismatch between InAs and Si. Quantitative measurement for threading dislocation density (TDD) in the InAs layers grown on GaAs and Si has been performed using transmission electron microscopy and electron channeling contrast imaging, both of which revealed that the TDD of InAs/Si sample is ∼35% higher than that of GaAs sample. Comparison of fabricated InAs p-i-n photodetectors indicated that reduction of threading dislocation density is crucial for low dark current and high responsivity mid-infrared photodetectors on Si.
Article
Full-text available
On-chip twisted light emitters are essential components of orbital angular momentum (OAM) communication devices1, 2. These devices address the growing demand for high-capacity communication systems by providing an additional degree of freedom for wavelength/frequency division multiplexing (WDM/FDM). Although whispering-gallery-mode-enabled OAM emitters have been shown to possess some advantages3, 4, 5, such as compactness and phase accuracy, their inherent narrow bandwidths prevent them from being compatible with WDM/FDM techniques. Here, we demonstrate an ultra-broadband multiplexed OAM emitter that utilizes a novel joint path-resonance phase control concept. The emitter has a micron-sized radius and nanometer-sized features. Coaxial OAM beams are emitted across the entire telecommunication band from 1,450 to 1,650 nm. We applied the emitter to an OAM communication with a data rate of 1.2 Tbit/s assisted by 30-channel optical frequency combs (OFCs). The emitter provides a new solution to further increase capacity in the OFC communication scenario.
Article
Full-text available
Semiconductor broadband light emitters have emerged as ideal and vital light sources for a range of biomedical sensing/imaging applications, especially for optical coherence tomography systems. Although near-infrared broadband light emitters have found increasingly wide utilization in these imaging applications, the requirement to simultaneously achieve both a high spectral bandwidth and output power is still challenging for such devices. Owing to the relatively weak amplified spontaneous emission, as a consequence of the very short non-radiative carrier lifetime of the inter-subband transitions in quantum cascade structures, it is even more challenging to obtain desirable mid-infrared broadband light emitters. There have been great efforts in the past 20 years to pursue high-efficiency broadband optical gain and very low reflectivity in waveguide structures, which are two key factors determining the performance of broadband light emitters. Here we describe the realization of a high continuous wave light power of >20 mW and broadband width of >130 nm with near-infrared broadband light emitters and the first mid-infrared broadband light emitters operating under continuous wave mode at room temperature by employing a modulation p-doped InGaAs/GaAs quantum dot active region with a ‘J’-shape ridge waveguide structure and a quantum cascade active region with a dual-end analogous monolithic integrated tapered waveguide structure, respectively. This work is of great importance to improve the performance of existing near-infrared optical coherence tomography systems and describes a major advance toward reliable and cost-effective mid-infrared imaging and sensing systems, which do not presently exist due to the lack of appropriate low-coherence mid-infrared semiconductor broadband light sources.
Article
Full-text available
Efficient complementary metal-oxide semiconductor-based nonlinear optical devices in the near-infrared are in strong demand. Due to two-photon absorption in silicon, however, much nonlinear research is shifting towards unconventional photonics platforms. In this work, we demonstrate the generation of an octave-spanning coherent supercontinuum in a silicon waveguide covering the spectral region from the near- to shortwave-infrared. With input pulses of 18 pJ in energy, the generated signal spans the wavelength range from the edge of the silicon transmission window, approximately 1.06 to beyond 2.4 μm, with a −20 dB bandwidth covering 1.124–2.4 μm. An octave-spanning supercontinuum was also observed at the energy levels as low as 4 pJ (−35 dB bandwidth). We also measured the coherence over an octave, obtaining , in good agreement with the simulations. In addition, we demonstrate optimization of the third-order dispersion of the waveguide to strengthen the dispersive wave and discuss the advantage of having a soliton at the long wavelength edge of an octave-spanning signal for nonlinear applications. This research paves the way for applications, such as chip-scale precision spectroscopy, optical coherence tomography, optical frequency metrology, frequency synthesis and wide-band wavelength division multiplexing in the telecom window.
Article
Full-text available
Background The present study is an attempt to explore the association between kitchen indoor air pollutants and physiological profiles in kitchen workers with microalbuminuria (MAU) in north India (Lucknow) and south India (Coimbatore). Methods The subjects comprised 145 control subjects, 233 kitchen workers from north India and 186 kitchen workers from south India. Information related to the personal and occupational history and health of the subjects at both locations were collected using a custom-made questionnaire. Worker lung function was measured using a spirometer. Blood pressure was monitored using a sphygmomanometer. Urinary MAU was measured using a urine analyzer. Indoor air monitoring in kitchens for particulate matter (PM), total volatile organic compounds (TVOC), carbon dioxide (CO2) and carbon monoxide (CO) was conducted using indoor air quality monitors. The size and shape of PM in indoor air was assessed using a scanning electron microscope (SEM). Fourier transform infrared (FTIR) spectroscopy was used to detect organic or inorganic compounds in the air samples. Results Particulate matter concentrations (PM2.5 and PM1) were significantly higher in both north and south Indian kitchens than in non-kitchen areas. The concentrations of TVOC, CO and CO2 were higher in the kitchens of north and south India than in the control locations (non-kitchen areas). Coarse, fine and ultrafine particles and several elements were also detected in kitchens in both locations by SEM and elemental analysis. The FTIR spectra of kitchen indoor air at both locations show the presence of organic chemicals. Significant declines in systolic blood pressure and lung function were observed in the kitchen workers with MAU at both locations compared to those of the control subjects. A higher prevalence of obstruction cases with MAU was observed among the workers in the southern region than in the controls (p < 0.01). Conclusions Kitchen workers in south India have lower lung capacities and a greater risk of obstructive and restrictive abnormalities than their north Indian counterparts. The study showed that occupational exposure to multiple kitchen indoor air pollutants (ultrafine particles, PM2.5, PM1, TVOC, CO, CO2) and FTIR-derived compounds can be associated with a decline in lung function (restrictive and obstructive patterns) in kitchen workers with microalbuminuria. Further studies in different geographical locations in India among kitchen workers on a wider scale are required to validate the present findings.
Article
Full-text available
Propagation of light in multimode optical fibers usually gives a spatial and temporal randomization of the transmitted field similar to the propagation through scattering media. Randomization still applies when scattering or multimode propagation occurs in gain media. We demonstrate that appropriate structuration of the input beam wavefront can shape the light amplified by a rare-earth-doped multimode fiber. Profiling of the wavefront was achieved by a deformable mirror in combination with an iterative optimization process. We present experimental results and simulations showing the shaping of a single sharp spot at different places in the output cross-section of an ytterbium-doped fiber amplifier. Cleaning and narrowing of the amplifier far-field pattern was realized as well. Tailoring the wavefront to shape the amplified light can also serve to improve the effective gain. The shaping approach still works under gain saturation, showing the robustness of the method. Modeling and experiments attest that the shaping is effective even with a highly multimode fiber amplifier carrying up to 127 modes.
Article
Short Wave Infra-Red (SWIR) photodetectors operating above the response cutoff of InGaAs- based detectors (1.7 - 2.5 μm) are required for both defense and civil applications. Type II Super-Lattices (T2SL) were recently proposed For near- room temperature SWIR detection as a possible system enabling bandgap adjustment in the required range. The work presented here focuses on a T2SL with alternating nano-layers of InGaAs and GaAsSb lattice- matched to an InP substrate. A near room temperature SWIR cutoff of 2.4 µm was measured. Electrical junctions were realized using Zn diffusion p-doping process. We realized and studied both mesa- and selective diffusion- based p-i-n photodiodes. Dark currents of mesa- based devices were 1.5 mA/cm² and 32 μA/cm² at 300 and 230 K respectively. Dark currents were reduced to 1.2 mA/cm² and 12 μA/cm² respectively by utilizing the selective diffusion process. The effect of operating voltage is discussed. At 300K the quantum efficiency was up to 40% at 2.18 µm in mesa devices. D∗ was at 2 μm.
Article
Abstract—We report a mid-infrared InAs p-i-n photodetector grown on GaAs using an abrupt method through cation exchange at the InAs/GaAs interface which has resulted in an extremely thin buffer layer. The InAs p-i-n photodetector has a photoresponsivity of 0.327 A/W at $\lambda$ =2.71 $\mu$ m which was comparable to the mid-infrared photodetector grown using a compositionally linear-graded process. This result showed that the abrupt method can be used to achieve mid-infrared photodetection on GaAs without the need of a compositionally graded buffer layer.
Article
We report a mid-infrared InAs p-i-n photovoltaic detector on a Si substrate operating at room temperature with a cutoff wavelength at 3.5 μm. The large 11.6% lattice mismatch between the device layer (InAs) and the Si substrate is accommodated by employing a compositionally graded buffer layer technique using solid-source molecular beam epitaxy and a commercially available GeSi substrate. An Al <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">x</sub> In <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">1-x</sub> As graded buffer layer is used to linearly scale the lattice constant from AlAs to InAs with an increasing x value. The photovoltaic detector has a responsivity of 0.57 A/W, measured using a blackbody source maintained at 700 °C.
Article
Si-based integrated circuits have been intensively developed over the past several decades through ultimate device scaling. However, the Si technology has reached the physical limitations of the scaling. These limitations have fuelled the search for alternative active materials (for transistors) and the introduction of optical interconnects (called “Si photonics”). A series of attempts to circumvent the Si technology limits are based on the use of III-V compound semiconductor due to their superior benefits, such as high electron mobility and direct bandgap. To use their physical properties on a Si platform, the formation of high-quality III-V films on the Si (III-V/Si) is the basic technology ; however, implementing this technology using a high-throughput process is not easy. Here, we report new concepts for an ultra-high-throughput heterogeneous integration of high-quality III-V films on the Si using the wafer bonding and epitaxial lift off (ELO) technique. We describe the ultra-fast ELO and also the re-use of the III-V donor wafer after III-V/Si formation. These approaches provide an ultra-high-throughput fabrication of III-V/Si substrates with a high-quality film, which leads to a dramatic cost reduction. As proof-of-concept devices, this paper demonstrates GaAs-based high electron mobility transistors (HEMTs), solar cells, and hetero-junction phototransistors on Si substrates.