ArticlePDF Available

Deep traps in GaAs/InGaAs quantum wells and quantum dots, studied by noise spectroscopy

Authors:

Abstract and Figures

Remotely doped In <sub>0.35</sub> Ga <sub>0.65</sub> As layers of different coverages 6, 9, 11, and 13 ML were grown by molecular beam epitaxy on (100) GaAs. Quantum dot (QD) nucleation was observed in situ by reflection high-energy electron diffraction at 8 ML growth of In <sub>0.35</sub> Ga <sub>0.65</sub> As , while for 6 ML, only two-dimensional (2D) growth was observed. Atomic force microscopy, low temperature photoluminescence, and Hall effect measurements confirmed this transition from 2D to three-dimensional growth. Low-frequency noise studies have been performed to probe defects in such heterostructures throughout the transition from a highly strained quantum well to QDs. Results were compared to a bulk n -type GaAs reference sample. We revealed three main defects in GaAs with activation energies of 0.8, 0.54, and 0.35 eV. These defects with the same activation energies were found in all samples. However, structures containing In <sub>0.35</sub> Ga <sub>0.65</sub> As QDs show an additional peak at low temperatures due to the presence of defects which are not observed for reference GaAs and quantum well samples. Detailed analysis shows that for 9 and 11 ML In <sub>0.35</sub> Ga <sub>0.65</sub> As QD samples this peak corresponds to the well known M1 defect in GaAs with an activation energy of 0.18 eV, while for a coverage of 13 ML the defect was found to have an activation energy of 0.12 eV. All defects were characterized quantitatively in terms of their activation energy, capture cross section, and density. These studies indicate that noise spectroscopy is a very sensitive tool for electronic material characterization on the nanoscale.
Content may be subject to copyright.
Deep traps in GaAs/InGaAs quantum wells and quantum dots,
studied by noise spectroscopy
Vas. P. Kunets,aT. Al. Morgan, Yu. I. Mazur, V. G. Dorogan, P. M. Lytvyn,bM. E. Ware,
D. Guzun, J. L. Shultz, and G. J. Salamo
Arkansas Institute for Nanoscale Materials Science and Engineering, University of Arkansas,
Fayetteville, Arkansas 72701, USA
Received 7 August 2008; accepted 26 September 2008; published online 19 November 2008
Remotely doped In0.35Ga0.65As layers of different coverages 6, 9, 11, and 13 ML were grown by
molecular beam epitaxy on 100GaAs. Quantum dot QDnucleation was observed in situ by
reflection high-energy electron diffraction at 8 ML growth of In0.35Ga0.65As, while for 6 ML, only
two-dimensional 2Dgrowth was observed. Atomic force microscopy, low temperature
photoluminescence, and Hall effect measurements confirmed this transition from 2D to
three-dimensional growth. Low-frequency noise studies have been performed to probe defects in
such heterostructures throughout the transition from a highly strained quantum well to QDs. Results
were compared to a bulk n-type GaAs reference sample. We revealed three main defects in GaAs
with activation energies of 0.8, 0.54, and 0.35 eV. These defects with the same activation energies
were found in all samples. However, structures containing In0.35Ga0.65As QDs show an additional
peak at low temperatures due to the presence of defects which are not observed for reference GaAs
and quantum well samples. Detailed analysis shows that for 9 and 11 ML In0.35Ga0.65As QD samples
this peak corresponds to the well known M1 defect in GaAs with an activation energy of 0.18 eV,
while for a coverage of 13 ML the defect was found to have an activation energy of 0.12 eV. All
defects were characterized quantitatively in terms of their activation energy, capture cross section,
and density. These studies indicate that noise spectroscopy is a very sensitive tool for electronic
material characterization on the nanoscale. © 2008 American Institute of Physics.
DOI: 10.1063/1.3020532
I. INTRODUCTION
During past decade, strain-driven epitaxy of III-V semi-
conductor QDsquantum dots have attracted the attention of
many researchers due to the unique physical properties of
QDs as a zero-dimensional quantum confined system and the
variety of applications in electronic and optoelectronic de-
vices. Indeed, fabrication of tunable, high efficient QD
lasers,1,2single or multicolored QD photodetectors35proved
their importance. There are an enormous number of scientific
reports on the skilful growth of well ordered, uniform, and
controlled density InAs QDs or InGaAs QDs of different
indium compositions. The optical properties of those were
studied extensively. Magnetotransport properties of QD
based heterostructures were studied less intensively and still
are of the great interest, especially for lateral transport. Most
of these reports were focused not only on the pure magne-
totransport properties of QDs themselves but also on elec-
tronic properties of high mobility channels remotely doped
quantum wellsin the presence of QDs.6,7Besides optical
and transport properties of zero-dimensional systems, their
noise properties, and the noise properties of heterosystems
where QDs are utilized as a main functional element are of
the great importance. Better knowledge of the noise proper-
ties would result in improved performance of QD based de-
vices and in understanding of physical limitations dictated by
nature that are difficult to overcome. Recently, the noise
properties of InAs QDs were reported for heterostructures
based on vertical transport.8,9Besides, 1/fflicker,
generation-recombination gr, and thermal noise, shot
noise in the InAs QD heterostructures was observed at very
low temperatures as a result of single electron tunneling
through a layer of QDs.10
Even though the formation of InAs or InGaAs self-
organized QDs is well understood by efficient manipulation
of the strain, there are still open questions about defect for-
mation within the volume of the QDs and in the layer cov-
ering them. Even though, current technology in molecular
beam epitaxy MBEallows for the growth of extremely
high quality dislocation-free QDs, these still need to be
probed on the electronic level. Some defects, such as vacan-
cies, cannot be detected by structural or morphological stud-
ies but can be revealed electronically. In addition, one has to
take into account that QDs are nanoscale objects and tech-
niques allowing for the probing of defects inside of them
must be extremely sensitive.
There were many successful attempts to investigate the
presence of QDs and defects in these QDs or adjusting layers
through space charge spectroscopies. For example, InAs/
GaAs QDs were studied by capacitance-voltage C-V
spectroscopy,11,12 and InAs/GaAs QDs, InP/GaInP QDs, and
In0.5Ga0.5As/GaAs QDs by deep level transient spectroscopy
DLTSand admittance spectroscopies.1315 In the present
aElectronic mail: vkunets@uark.edu.
bPermanent address: V. Lashkaryov Institute of Semiconductor Physics,
NAS of Ukraine, Kiev 03028, Ukraine 1.
JOURNAL OF APPLIED PHYSICS 104, 103709 2008
0021-8979/2008/10410/103709/8/$23.00 © 2008 American Institute of Physics104, 103709-1
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
work we are using deep level noise spectroscopy DLNSto
probe defects caused by In0.35Ga0.65As QDs grown between
GaAs barriers. We show that at low temperatures additional
defects appear depending on the In0.35Ga0.65As coverage.
These are a 0.18 eV deep trap for 9 and 11 ML and a 0.12 eV
deep trap for 13 ML coverage. No such defects are revealed
in the same temperature range for reference GaAs and In-
GaAs QW samples. In addition we observe narrowing of full
width at half maximum FWHMfor 0.54 and 0.35 eV re-
lated defects typical for MBE grown GaAs with insertion of
In0.35Ga0.65As layers resulting in better resolution of noise
peaks. The possible mechanisms of this enhancement are dis-
cussed. These results are in good agreement with results of
Fang et al.14 observed by the technique of DLTS.
II. EXPERIMENTAL DETAILS
Using a Riber 32-P solid-source MBE system, we grew
eight different samples on semi-insulating 100oriented
GaAs wafers. After removing the oxide layer at 610 ° C in an
As4flux, an undoped 5000 Å thick GaAs buffer layer was
grown at a rate of 1 ML/s. Then 5000 Å of uniformly Si-
doped GaAs Nd=71016 cm−3was grown at a substrate
temperature of 580 °C. A 200 Å thick undoped GaAs spacer
layer was then grown at the same temperature. At this point
growth was stopped and the substrate temperature was de-
creased to 540 °C, where a 200 Å thick undoped GaAs
spacer was grown. Then the substrate temperature was raised
back to 580 °C and the rest of the 1500 Å uniformly Si-
doped GaAs Nd=71016 cm−3cap layer was grown. The
growth conditions for next four samples were the same as for
this reference sample, S0, except that 6 ML sample S6,9
ML sample S9,11MLsample S11, and 13 ML sample
S13of In0.35Ga0.65As were grown following the initial
200 Å spacer. The growth temperature for the InGaAs layers
was kept at 540 °C. The growth was monitored by reflection
high-energy electron diffraction RHEEDand the surface
temperature was monitored by band edge measurement using
transmitted light in situ. The nucleation of QDs was observed
at about 8 ML of In0.35Ga0.65As growth. At the end we grew
three additional uncapped QD samples with 9, 11, and 13
ML of In0.35Ga0.65As for atomic force microscopy AFM
studies. The substrate temperature was rapidly cooled down
after each of the growths of the uncapped QDs.
The photoluminescence PLmeasurements were per-
formed at temperature of 10 K in a closed-cycle helium cry-
ostat. The excitation was done with a frequency-doubled yt-
trium aluminum garnet laser at 532 nm. The laser spot
diameter was 20
m and the excitation power density was
20 W/cm2.
Using standard optical photolithography, four-terminal
Greek-cross structures, with square shaped active area of
2020
m2were fabricated. Chemical wet etching was
used for mesa device isolation. Ohmic contacts were formed
by alloying an evaporated AuGe/Ni/Au 75/20/250 nm
metal films. Ohmic contact resistance was optimized using a
fast thermal ramp up to 420 ° C followed by a 2 min anneal-
ing.
The van der Pauw technique was used for resistivity and
Hall effect measurements using a benchtop electromagnet
from MMR technologies at a magnetic field of 0.25 T. The
transverse noise was measured between Hall terminals using
an SR560 differential low-noise preamplifier and a SR785
noise spectrum analyzer in a well-shielded environment on
vibration free table. The frequency range of our measure-
ments covered 3 Hz to 100 kHz. The samples were biased by
a low-noise battery pack 12/24 Vin series with a low-noise
metal film load resistor RL. The current in the circuit can be
changed by choosing RL, where RLRsample. The Hall effect
and noise measurements were made in temperature range of
80–400 K.
III. RESULTS AND DISCUSSION
AFM images of the open QD samples with different
In0.35Ga0.65As coverages are shown in Figs. 1a1c. Our
structural analysis shows that height, density, and size distri-
bution depend on the number of deposited In0.35Ga0.65As
monolayers. For all coverages we observe only single mode
size distribution. For 9 ML of In0.35Ga0.65As we observe QDs
with an average height of 34 Å and areal density of NQD
=3.81010 cm−2. For 11 ML coverage, the height and areal
density of QDs increase to 47 Å and 8.41010 cm−2, re-
spectively. Further increase in coverage to 13 ML resulted in
a narrower size distribution Fig. 1c, a lower density of
QDs NQD=7.21010 cm−2 compared to 11 ML coverage,
and the height being of 54 Å. It is necessary to mention that
a)
b
)
c) 13 ML
02468
0
20
40
60
80
Count
11 ML
02468
0
20
40
60
Count
9ML
02468
0
20
40
60
Count
Hei
g
ht
(
nm
)
FIG. 1. Color onlineAFM images of In0.35Ga0.65As QDs with different
coverages: a9MLsample S9,b11 ML sample S11,andc13 ML
sample S13. The height distributions are shown on the right. Enlarged 3D
images of QDs with different coverages are shown on the insets, where
lateral dimensions are 200200 nm2.
103709-2 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
shape of the dots is elliptical with an elongation along the
011
¯
crystallographic direction as was reported before.16
The PL measurements for covered samples are shown in
Fig. 2. There is a noticeable redshift of the PL line with
increasing coverage. The maxima of the PL lines are mea-
sured at 1.301, 1.249, and 1.207 eV for 9, 11, and 13 ML,
respectively. Each PL line is well fit by a single Gaussian.
This indicates a single size distribution of the capped QDs.
The PL results are in a very good agreement with size dis-
tribution analysis shown in Figs. 1a1c. From the Gauss-
ian fit of each PL band, the FWHM was determined. We
have measured FWHM of 38, 48, and 45 meV for 9, 11, and
13 ML, respectively. The lower FWHM for the 13 ML
sample agrees with AFM statistics, where the size distribu-
tion for the 13 ML sample also exhibits a smaller FWHM.
Another important feature is the PL line intensity as a func-
tion of coverage. Assuming defect-free QDs with a single
size distribution, one would expect higher integral PL inten-
sity for the samples with higher QD density. However, we
observe the opposite behavior, i.e., increasing the coverage
leads to a decrease in the integral PL intensity even though
the QD density does increase. We assume that such behavior
could be caused by appearing of additional channels of non-
recombinative radiation with coverage increase. This is a
topic of our further discussions.
The Hall effect measurements were performed to study
conductivity in the wide 80–400 Ktemperature range and to
provide required information for noise data analysis. The
Hall mobilities as function of temperature and coverage are
shown in Fig. 3. The reference GaAs sample S0exhibits a
typical mobility curve versus temperature.17 At high tem-
peratures the absolute value of mobility is limited due to
phonon scattering. By lowering the lattice temperature this
scattering mechanism becomes less efficient leading to an
increase in mobility. At low enough temperatures, scattering
by ionized impurities dominates and mobility is driven
mostly by impurity scattering. The resulting mobility peaks
at temperatures below 150 K and then decreases. In contrast
to the reference GaAs sample, sample S6 consisting of 6 ML
of In0.35Ga0.65As shows behavior typical for a remotely
doped quantum well. At high temperatures the mobility is
about the same as that of bulk GaAs. This is expected be-
cause at high temperatures the InGaAs quantum well is too
narrow to confine the conduction electrons only within In-
GaAs quantum well, and conductivity through the GaAs bar-
riers takes a place. However, at low temperatures the con-
ductivity mostly occurs through the InGaAs channel and less
through the Si-doped GaAs barriers. Thus, for this tempera-
ture range the mobility does not show a peak as in the case of
GaAs, because the electron gas is localized two dimension-
ally in the QW and experiences less scattering from ionized
donors in the barrier. However the absolute value of mobility
is not so high and conductivity in adjacent GaAs layers is
contributing. The temperature behavior of the mobility re-
veals the two-dimensional 2Dnature of the growth for 6
ML of In0.35Ga0.65As, which agrees with in situ RHEED ob-
servations. Mobility curves for samples with QDs are similar
to the reference GaAs sample. However the measured values
are lower. Besides, at low temperatures it is obvious that
samples S11 11 MLand S13 13 MLwith higher densities
of QDs show lower mobility than the sample S9 9ML.
This result is expected and can be explained if one will con-
sider each QD not only as potential quantum well were an
electron can spend some time but also as a localized charged
center with an ability to scatter the majority of carriers.18
Having the necessary information about the structural
and electronic properties of the samples, we studied their
noise characteristics. As an example, the room temperature
noise spectrum of the reference GaAs sample is shown in
Fig. 4. The noise voltage spectral density versus frequency
shows complex behavior. Besides 1/fflickernoise and
thermal noise, the spectrum is distorted by two additional
Lorentzians at low and high frequencies due to the presence
of local centers in the band gap of the semiconductor. The fit
of the experimental noise spectrum can be done according to
SV,noise =B
f+
i
Ai
1
1+2
f
i2+4kBTR,1
where Band Aiare the amplitudes of the 1/fand grpro-
cesses in the volume of the semiconductor, fis the fre-
1.10 1.15 1.20 1.25 1.30 1.35 1.4
0
13 ML
11 ML
PL
S
ignal
(
arb. units
)
Photon Energy (eV)
9ML
T=10K
Iex =20W/cm
2
FIG. 2. PL spectra of In0.35Ga0.65As QDs with different coverages of 9 ML
sample S9,11MLsample S11, and 13 ML sample S13measured at
T=10 K and excitation power of 20 W /cm2.
50 100 150 200 250 300 350 400
4x103
5x103
6x103
7x103
8x103
9x103
11 ML
9ML
13 ML
0ML
6ML
Mobility (cm
2
/Vs)
Temperature
(
K
)
FIG. 3. Temperature dependent measurements of electron mobility for the
given coverages of In0.35Ga0.65As. Here a coverage of 0 ML represents the
reference GaAs sample.
103709-3 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
quency,
iis the characteristic time constant of the grpro-
cess, kBis the Boltzmann constant, Tis the temperature, and
Ris the sample resistance. Fitting the data in Fig. 4to Eq.
1, we found that grprocesses are characterized by
1
=4 ms and
2=24
s time constants.
By measuring the temperature dependence of
iand by
plotting ln
iversus 1/T, one can determine the parameters
of the local levels in the band gap causing grprocesses in
the semiconductor. Another method, which has an excellent
theoretical background for the analysis of the local centers in
semiconductors, was proposed in Ref. 19. We will follow
these techniques to characterize local centers in our samples.
Figure 5shows the relative noise spectral density S
=SV/V2versus temperature for a frequency of 20 Hz for each
of the samples in this study. Here, Vis the voltage drop
across the sample. These data were taken from noise spectra
similar to that shown in Fig. 4measured through a tempera-
ture range of 80–400 K at steps of 10 K. For the reference
GaAs sample, S0, there are three peaks. These peaks, labeled
A, B, and C, are more clearly found in a linear scale, which
is shown in the inset in Fig. 5. The insertion of an
In0.35Ga0.65As layer into the GaAs leads to a pronounced
sharpening of the peaks. A similar behavior was observed by
Fang et al.,14 who applied DLTS for studies of In0.5Ga0.5As
QDs embedded in GaAs. In addition to this observation for
QD systems, we show here that even fora6MLchannel of
In0.35Ga0.65As where there are no QDs we see these changes
clearly. This can be a good indicator that the sensitivity of
DLNS is good enough for studies on nanoscale materials. As
a new observation, we find that for the 9, 11, and 13 ML
samples which resulted in the formation of QDs an addi-
tional low temperature peak has appeared in the noise spec-
tra. We could not detect this peak for reference sample or the
6MLIn
0.35Ga0.65As QW sample. Thus, we associate this
peak, D, uniquely with the presence of QDs in the GaAs
matrix.
To prove that peaks A, B, C, and D are defect related we
have performed studies of them as a function of temperature
and frequency. The peaks related to the local levels in semi-
conductors shift toward higher temperatures as well as de-
crease in intensity as the frequency is increased.1921 Such
tendency is represented in Fig. 6for the 11 ML sample.
Noise data for other samples show similar behavior as func-
tion of temperature and frequency.
Following the work done in Ref. 19, and assuming that
10010110210310410
5
g-r
g-r
Johnson Noise, 4kBTR
1/f
i= 893.6 A
T = 294 K
S0
10-
12
10-13
10-14
10-15
10-16
10-17
SV(V
2
/Hz)
f
(
Hz
)
FIG. 4. Noise spectrum from the reference GaAs sample measured at bias
current of 893.6
AandT=294 K across the Hall terminals of the Greek-
cross structure. The fitting was done according to Eq. 1. Each spectral
component participating in conductivity fluctuations is shown separately.
100 150 200 250 300 350 400 45
0
D
C
B
AS13
S11
S9
S6
S0
SV/V
2(arb. units)
Temperature
(
K
)
50 100 150 200 250 300 350 400
1.0x10-14
6.0x10-14
1.1x10-13
S0
C
B
A
Temperature (K)
SV/V2(Hz-1)
FIG. 5. Relative noise spectral density vs temperature for frequency of 20
Hz in a semilogarithmic scale. The curves are shown for all five samples: S0
reference GaAs; S6, S9, S11, and S13 are 6, 9, 11, and 13 ML of
In0.35Ga0.65As, respectively. The curves are shifted for clarity. The inset
shows the relative noise spectral density for sample S0 in a linear scale.
50 100 150 200 250 300 350 40
2560 Hz
1280 Hz
640 Hz
320 Hz
160 Hz
80 Hz
40 Hz
20 Hz
10 Hz
5Hz
S11 (11 ML)
D
C
B
A
10-
10-13
10-14
10-15
10-16
10-17
SV/V2(Hz-1)
Temperature (K)
FIG. 6. Relative noise spectral density measured as function of temperature
and frequency for 11 ML In0.35Ga0.65As QD sample. Four defects labeled as
A, B, C, and D are distinctly detected.
103709-4 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
ithe capture cross section,
, depends exponentially on
temperature,
=
0expE1/kBT, and iithe time constant
iassociated to the return to the equilibrium of the occupa-
tion of the level is given by
=
cF=
c0exp
E1
kBT
F,2
the relative noise spectral density can be written as19,22
S=SV
V2=A
F1−F
1+
2=A
c0expE1/kBTF21−F
1+
2
c0
2exp2E1/kBTF2,3
where
c0=
0
Tn−1 and Fis the occupancy of a level given
by F=1/1+exp关共EFE0/kBT兴兲. Here E1is the activation
energy, E0is the energy position of the deep level in the band
gap with respect to the bottom of conduction band, EFis the
Fermi energy,
cis the capture time constant,
Tis the ther-
mal velocity, nis the electron concentration,
=2
f, and
A=4Nt/VNd
2, where Ntis the trap density, Vis the sample
volume, and Ndis the concentration of the shallow donors.
To characterize the deep levels, we plotted dependencies
of lnSmaxon ln
and 1/kBTmax on ln
for each defect
and for all five samples. Figure 7shows these plots for QW
sample with 6 ML and QD sample with 11 ML of
In0.35Ga0.65As. The generalized results for all samples are
comprised in Table I.
The analysis of lnSmaxversus ln
for defect A de-
tected at high temperatures see Fig. 5shows that this de-
pendence can be very well fitted by a straight line with a
slope of 0.9. According to Ref. 19, we have EFlocated
above the trap level E0and E1E0. For this case, assuming
F21 and solving the extremum problem for Eq. 3, one
can find an activation energy E1and energy location of the
deep level E0as
E1=1
tan
T
,E0=1 − tan
s
tan
T
,4
where tan
Tand tan
sare the slopes of 1/kBTmax versus
ln
and lnSmaxversus ln
, respectively. We found ac-
tivation energy E1to be 0.8 eV and E0located at about
100 meV below the conduction band. For temperature range
of 300–360 K, where defect A is detected, the Fermi level for
GaAs is found to be 0.05 eV 300 Kand 0.07 eV 360 K
located below the conduction band. According to our analy-
sis we have not exactly the case E0EFkBT, however, the
Fermi level is still above the trap over the whole temperature
range. Thus, we used the equations proposed in Ref. 19 for
calculating the trap density, Nt, and capture cross section,
0.
For a frequency, f=320 Hz, measured electron density, n
=61016 cm−3,
T=4.4107cm/s, and density of states
Nc=51017 cm−3, we found
010−10–10−11 cm2and Nt
1014 cm−3.
Trap B was detected for all five samples in the tempera-
ture range of 240–320 K. Analysis of lnSmaxversus ln
gives an average slope close to 1. This situation corresponds
to the case when E1E0and is unfavorable for the interpre-
tation of deep levels in semiconductors by noise
spectroscopy.19,2325 Special techniques that inject minority
carriers band-to-band illumination or injection through a
forward-biased p-njunctionwere proposed to determine the
local level parameters.24,26,27 Following Refs. 25 and 28, one
can determine the activation energy, E1, using Eq. 4. For
defect B we find that the average activation energy for the
studied samples is 0.54 eV. Then the time constant
Tmaxis given by28
Tmax=
0exp
E1
kBT
,5
where the relation
Tmax=2
f
Tmax=1 6
holds true.28
Then
0can be calculated as28
0=
exp
E1
kBTmax
=1
n
T
0
=expE1/kBTmax
n
T
Tmax.7
For frequency fmax = 320 Hz, n=61016 cm−3,
T=4.4
107cm/s, and Smax310−15 Hz−1, we have
0in the
range of 10−12–10−11 cm2. The defect density can be deter-
46810
30
40
50
60
7
0
796 meV
552 meV
388 meV
C
B
A
1/kBT(eV-
1
)
Ln
46810
-38
-36
-34
-32
-30
-28
S6
C
B
A
~0.99
~0.9
~0.87
LnS
max
Ln
(
a)
246810
30
40
50
60
70
80
90
100
110
943 meV
544 meV
323 meV
176 meV
D
C
B
A
1/kBT(eV-
1
)
Ln
246810
-38
-36
-34
-32
-30
-28
S11
D
CB
~0.99
~1.05
~0.98
~0.89
A
LnS
max
Ln
(
b
)
FIG. 7. 1 /kBTmax vs ln
兲共the Arrhenius plotand lnSmaxon ln
for a
6 ML QW and b11 ML In0.35Ga0.65As QD samples. The activation ener-
gies left graphand slopes right graphare depicted for defects A, B, C,
and D.
103709-5 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
mined for this case only very roughly by using25
Nt=4
fmaxSmaxn2V.8
Assuming the occupancy of the local level to be about 0.5,
we have a trap density of defect B approximately 6
1013 cm−3.
Trap C was detected in all five samples in the tempera-
ture range of 170–210 K. Analysis of the relative noise den-
sity amplitude versus frequency shows that the slope is again
close to 1. Thus, for extracting information about this defect
we used Eqs. 48. We have that for defect C with an
average activation energy of E1=0.35 eV at fmax =80 Hz,
n=61016 cm−3,
T=4.4107cm/s, and Smax=5
10−16 Hz−1 the value of
0is in the range of
10−13–10−12 cm2. The trap density was estimated like in the
previous case and was found to be 31012 cm−3.
Peak D measured at low temperatures 100–140 Kis
the most interesting one observed in this work. Because traps
A, B, and C are present in all samples including the GaAs
reference, we can conclude that they are GaAs related
defects.14,29 Trap D, however, appears to be unique to only
the samples containing QDs as it is not found in the refer-
ence GaAs sample or the 6 ML In0.35Ga0.65As QW sample.
Thus, we assume that at low temperature peak D is caused by
defects related to growth of In0.35Ga0.65As QDs in GaAs and
not simply of strained 6 ML In0.35Ga0.65As 2D layer.
Analysis of lnSmaxversus ln
shows that slope of this
dependence is again close to 1 see Table Ifor samples S9,
S11, and S13 with QDs. The activation energy can be found
from Eq. 4. From an Arrhenius plots similar to one shown
in Figs. 7aand 7b, we found that the D trap activation
energies are 0.18, 0.18, and 0.12 eV for the 9, 11, and 13 ML
samples, respectively. Using n=61016 cm−3 measured by
Hall effect, and
T=4.4107cm/s, we estimated
0to be
110−15 cm2for samples S9 fmax=15 Hzand S11
fmax=20 Hz. For sample S13 with 13 ML of In0.35Ga0.65As
with an electron density of n=51016 cm−3 and
T=4.4
107cm/s, we have
0510−18 cm2. The defect densi-
ties can be roughly estimated by Eq. 8, where trap densities
of 71011 cm−3,11011 cm−3, and 81010 cm−3 were
found for samples with 9, 11, and 13 ML of In0.35Ga0.65As,
respectively. The large difference in the activation energies
E1and capture cross section
0allows us to assume that
peak D probably has a complex behavior as it was shown in
Ref. 14. It was revealed by DLTS Ref. 14that in the same
temperature range there are two overlapping peaks iM1
with activation energy of 0.18 eV which is a typical trap for
MBE-GaAs materials and iipeak D with activation energy
of 0.1 eV which is related to In0.5Ga0.5As QDs grown on
GaAs. The capture cross sections were found to be 1.6
10−15 cm2for M1 trap and 5.410−18 cm2for D trap. It
is obvious that values that we measured by noise spectros-
copy are in good agreement with values reported by DLTS.
The question is how activation energies of 0.18 and 0.12 eV
measured by noise can be correlated with DLTS values re-
ported in Ref. 14? Indeed, these two defects are observed
only for samples where QDs are present. Then, it is logical to
assume that these defects are located nearby or inside of
QDs. Their most probable location is the 200 Å thick un-
doped GaAs spacer layer. While the spacer layer is fully
TABLE I. Activation energy E1, energy of the local level E0, trap density Nt, and
0obtained after analysis of
the temperature dependence of noise spectra.
S0
Ref.
S6
6ML
S9
9ML
S11
11 ML
S13
13 ML
Defect A
Slope 0.86 0.87 0.90 0.89 0.90
E1eV0.75 0.79 0.75 0.94 0.76
E0eV0.108 0.103 0.075 0.104 0.077
0cm2410−11 410−10 410−11 310−8 710−11
Ntcm−33.11014 5.9 1014 1.11014 2.31014 9.2 1013
Defect B
Slope 1.04 0.90 1.07 0.98 0.94
E1eV0.53 0.55 0.57 0.54 0.52
0cm2110−12 310−12 810−12 410−12 110−12
Ntcm−37.51013 6.3 1013 5.91013 4.91013 5.4 1013
Defect C
slope 1.07 0.99 1.06 1.05 1.02
E1eV0.33 0.38 0.32 0.32 0.35
0cm2510−14 210−12 510−14 110−13 110−13
Ntcm−31.61012 3.2 1012 3.11012 2.41012
Defect D
Slope 1.09 0.99 0.96
E1eV0.18 0.18 0.12
0cm2810−16 110−15 510−18
Ntcm−36.61011 1.4 1011 8.11010
103709-6 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
depleted, we have a situation similar to that reported in Ref.
25 where the energy, E1, determined by Eq. 4gives the
level position E0responsible for the observed noise peak.
Then the time constant will be given by the Schockley–Hall–
Read theory and Eq. 5holds true, where
0will be given
by25
0=
0
TNc−1,9
and Ncis the density of states in the conduction band. Then
we have
0equal to 110−15 cm2for samples S9 and S11
and 510−18 cm2for sample S13. The densities of traps are
the same and are given by Eq. 8. However, according to
our analysis, we cannot assume that a trap with an energy of
0.12 eV is located in the InGaAs band gap. If this would be
the case, then the Fermi level is located significantly higher
than the trap level E0. Indeed, according to our simple esti-
mations based on the PL data and assuming the ratio of the
InGaAs band offset relative to GaAs to be Ec:Ev
=0.64:0.36, we would have the ground state located at about
165 meV below the GaAs band for 13 ML In0.35Ga0.65As in
approximation of quantum well, where the thickness of the
QW was equal to height of the QDs. For temperature range
of 100–140 K and free carrier density of 61016 cm−3,we
have the Fermi level located at 3 meV for 100 K and at 10
meV for 140 K below the GaAs conduction band. Compared
to the trap energy E0, the Fermi level is 282 meV above E0
for 100 K and 275 meV above E0for 140 K. The slope of
lnSmaxversus ln
dependence would be less than 1.
While experimental data show the slope to be close to 1, we
can only assume that the trap with activation energy of 0.12
eV is physically located in the nearby GaAs spacer and
caused by the presence of the QDs. Besides, if this defect
does not propagate deep into n-type GaAs cap layer than
density of the traps with activation energy of 0.12 eV could
be overestimated one has to take correction for the volume
V. The small value of
0for defects caused by growth of
In0.5Ga0.5As Ref. 14or InAs QDs was reported30 before
and agrees well with our results. Another trap which is re-
vealed in samples with 9 and 11 ML is probably the well
known M1 trap in MBE grown GaAs. The M1 trap with
activation energy of 0.18 eV is related to point defects in
GaAs.14,29 It is quite reasonable to assume that due to the
high lattice mismatch between InGaAs QDs and the GaAs
cap layer, the density of point defects in this region raises
becoming resolved by our technique. In addition, we observe
also a narrowing of peaks B and C. As result, these became
much better resolved by our noise technique. There are two
possible mechanisms of this enhancement ipresence of the
strain field due to the lattice mismatch and/or iichanges in
electron population in adjusting GaAs barrier layers due to
the presence of the quantum reservoirs such as a quantum
well or QDs. The depopulation of free carriers in GaAs bar-
riers could lead to lowering of the Fermi level and, thus, get
closer to resonance with the deep trap E0. Because we do not
see significant enhancement of the densities for defects B
and C with increasing number of In0.35Ga0.65As monolayers,
the second assumption is the most probable.
IV. CONCLUSIONS
Using noise spectroscopy and MBE grown
In0.35Ga0.65As layers of different thicknesses on GaAs, we
have studied deeps levels at the transition from 2D to three-
dimensional 3Dgrowth. The results were compared to a
reference n-type GaAs sample. We found three main defects
with activation energies of 0.8, 0.54, and 0.35 eV in refer-
ence GaAs. The same defects were detected for 6 ML
In0.35Ga0.65As highly strained quantum well as well as struc-
tures with QDs. At the transition to 3D, QD growth, one
additional defect appears with activation energy of 0.18 eV
which is a well known M1 trap level typical for n-type GaAs
grown by MBE and is most likely related to point defects.
Further increase in the InGaAs coverage above 11 ML leads
to an additional defect that is most probably located in GaAs
spacer layer. This trap is located at 0.12 eV below the GaAs
conduction band and has a small capture cross section. The
results are in agreement with the data received using DLTS
for self-organized In0.5Ga0.5As QDs grown on GaAs.14 This
work demonstrates that noise spectroscopy is a sensitive tool
with the capability to probe nanoscaled materials.
ACKNOWLEDGMENTS
This work was supported by the National Science Foun-
dation under Grant No. DMR-0520550.
1S. Fafard, K. Hinzer, S. Raymond, M. Dion, J. McCaffrey, Y. Feng, and S.
Charbonneau, Science 274,13501996.
2C. Y. Jin, H. Y. Liu, K. M. Groom, Q. Jiang, M. Hopkinson, T. J. Badcock,
R. J. Royce, and D. J. Mowbray, Phys. Rev. B 76, 085315 2007.
3H. C. Liu, M. Gao, J. McCaffrey, Z. R. Wasilewski, and S. Fafard, Appl.
Phys. Lett. 78,792001.
4S. Maimon, E. Finkman, G. Bahir, S. E. Schacham, J. M. Garcia, and P.
M. Petroff, Appl. Phys. Lett. 73, 2003 1998.
5Z. Ye, J. C. Campbell, Z. Chen, E. T. Kim, and A. Madhukar, J. Appl.
Phys. 92, 4141 2002.
6H. Sakaki, G. Yusa, T. Someya, Y. Ohno, T. Noda, H. Akiyama, Y. Ka-
doya, and H. Noge, Appl. Phys. Lett. 67,34441995.
7I. R. Pagnossin, E. C. F. da Silva, A. A. Quivy, S. Martini, and C. S.
Sergio, J. Appl. Phys. 97, 113709 2005.
8A. Tsormpatzoglou, N. A. Hastas, D. H. Tassis, C. A. Dimitriadis, G.
Kamarinos, P. Frigeri, S. Franchi, E. Gombia, and R. Mosca, Appl. Phys.
Lett. 87, 163109 2005.
9N. Arpatzanis, A. Tsormpatzoglou, C. A. Dimitriadis, J. D. Song, W. J.
Choi, J. I. Lee, and C. Charitidis, J. Appl. Phys. 102, 054302 2007.
10P. Barthold, F. Hohls, N. Maire, K. Pierz, and R. J. Haug, Phys. Rev. Lett.
96, 246804 2006.
11P. N. Brounkov, A. Polimeni, S. T. Stoddart, M. Henini, L. Eaves, P. C.
Main, A. R. Kovsh, Yu. G. Musikhin, and S. G. Konnikov, Appl. Phys.
Lett. 73, 1092 1998.
12G. Medeiros-Ribeiro, D. Leonard, and P. M. Petroff, Appl. Phys. Lett. 66,
1767 1995.
13S. Anand, N. Carlsson, M. E. Pistol, L. Samuelson, and W. Seifert, J.
Appl. Phys. 84, 3747 1998.
14Z. Q. Fang, Q. H. Xie, D. C. Look, J. Ehret, and J. E. van Nostrand, J.
Electron. Mater. 28,L131999.
15P. Krispin, J. L. Lazzari, and H. Kostial, J. Appl. Phys. 84,61351998.
16S. O. Cho, Zh. M. Wang, and G. J. Salamo, Appl. Phys. Lett. 86, 113106
2005.
17G. E. Stillman and C. M. Wolfe, Thin Solid Films 31,691976.
18D. C. Look, Electrical Characterization of GaAs Materials and Devices
Wiley, New York, 1989, p. 280.
19M. E. Levinshtein and S. L. Rumyantsev, Semicond. Sci. Technol. 9,1183
1994.
20D. D. Carey, S. T. Stoddart, S. J. Bending, J. J. Harris, and C. T. Foxon,
Phys. Rev. B 54, 2813 1996.
21M. J. Deen, M. E. Levinshtein, S. L. Rumyantsev, and J. Orchard-Webb,
103709-7 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
Semicond. Sci. Technol. 14, 298 1999.
22J. A. Copeland, IEEE Trans. Electron Devices 18,501971.
23M. E. Levinshtein, S. L. Rumyantsev, J. W. Palmour, and D. B. Slater, Jr.,
J. Appl. Phys. 81, 1758 1997.
24J. W. Palmour, M. E. Levinshtein, and S. L. Rumyantsev, Semicond. Sci.
Technol. 11,11461996.
25S. L. Rumyantsev, N. Pala, M. S. Shur, E. Borovitskaya, A. P. Dmitriev,
M. E. Levinshtein, R. Gaska, M. A. Khan, J. Yang, X. Hu, and G. Simin,
IEEE Trans. Electron Devices 48, 530 2001.
26N. V. Dyakonova, M. E. Levinshtein, and S. L. Rumyantsev, Sov. Phys.
Semicond. 25, 1241 1991.
27M. E. Levinshtein, J. W. Palmour, and S. L. Rumyantsev, Semicond. Sci.
Technol. 9, 2080 1994.
28N. V. Dyakonova, M. E. Levinshtein, and S. L. Rumyantsev, Semicond.
Sci. Technol. 11,1771996.
29N. Chand, A. M. Sergent, J. P. van der Ziel, and D. V. Lang, J. Vac. Sci.
Technol. B 7,3991989.
30M. M. Sobolev, A. R. Kovsh, V. M. Ustinov, A. Yu. Egorov, A. E.
Zhukov, M. V. Maksimov, and N. N. Ledenstsov, Semiconductors 31,
1074 1997.
103709-8 Kunets et al. J. Appl. Phys. 104, 103709 2008
Downloaded 26 Nov 2009 to 217.20.172.209. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
... Also, Fang et al [34] detected the M1 defect in In 0.5 Ga 0.5 As/GaAs QDs structures grown by MBE and they attributed this defect to point defects instead of defect-impurity complexes. Moreover, Kunets et al [48] observed the M1 trap in In 0.35 Ga 0.65 As/GaAs QD structures grown by MBE using noise spectroscopy measurements, and they related the increase of its density in the vicinity of In 0.35 Ga 0.65 As QDs to strain. Thus there is a consensus that E2 PIN and E3 QWR are related to M1 defect, which could be assigned to defect-impurity complexes or/and point defects [34,45,48,49]. ...
... Moreover, Kunets et al [48] observed the M1 trap in In 0.35 Ga 0.65 As/GaAs QD structures grown by MBE using noise spectroscopy measurements, and they related the increase of its density in the vicinity of In 0.35 Ga 0.65 As QDs to strain. Thus there is a consensus that E2 PIN and E3 QWR are related to M1 defect, which could be assigned to defect-impurity complexes or/and point defects [34,45,48,49]. The shallow trap E1 QWR with energy of ∼10 meV is only observed in QWR undoped devices. ...
Article
Full-text available
InGaAs quantum wire (QWr) intermediate-band solar cell–based nanostructures grown by molecular beam epitaxy are studied. The electrical and interface properties of these solar cell devices, as determined by current–voltage (I–V) and capacitance–voltage (C-V) techniques, were found to change with temperature over a wide range of 20–340 K. The electron and hole traps present in these devices have been investigated using deep-level transient spectroscopy (DLTS). The DLTS results showed that the traps detected in the QWr-doped devices are directly or indirectly related to the insertion of the Si δ-layer used to dope the wires. In addition, in the QWr-doped devices, the decrease of the solar conversion efficiencies at low temperatures and the associated decrease of the integrated external quantum efficiency through InGaAs could be attributed to detected traps E1QWR_D, E2QWR_D, and E3QWR_D with activation energies of 0.0037, 0.0053, and 0.041 eV, respectively.
... Also, Fang et al [34] detected the M1 defect in In 0.5 Ga 0.5 As/GaAs QDs structures grown by MBE and they attributed this defect to point defects instead of defect-impurity complexes. Moreover, Kunets et al [48] observed the M1 trap in In 0.35 Ga 0.65 As/GaAs QD structures grown by MBE using noise spectroscopy measurements, and they related the increase of its density in the vicinity of In 0.35 Ga 0.65 As QDs to strain. Thus there is a consensus that E2 PIN and E3 QWR are related to M1 defect, which could be assigned to defect-impurity complexes or/and point defects [34,45,48,49]. ...
... Moreover, Kunets et al [48] observed the M1 trap in In 0.35 Ga 0.65 As/GaAs QD structures grown by MBE using noise spectroscopy measurements, and they related the increase of its density in the vicinity of In 0.35 Ga 0.65 As QDs to strain. Thus there is a consensus that E2 PIN and E3 QWR are related to M1 defect, which could be assigned to defect-impurity complexes or/and point defects [34,45,48,49]. The shallow trap E1 QWR with energy of ∼10 meV is only observed in QWR undoped devices. ...
Article
InGaAs quantum wire (QWr) intermediate-band solar cell–based nanostructures grown by molecular beam epitaxy are studied. The electrical and interface properties of these solar cell devices, as determined by current–voltage (I–V) and capacitance–voltage (C-V) techniques, were found to change with temperature over a wide range of 20–340 K. The electron and hole traps present in these devices have been investigated using deep-level transient spectroscopy (DLTS). The DLTS results showed that the traps detected in the QWr-doped devices are directly or indirectly related to the insertion of the Si δ-layer used to dope the wires. In addition, in the QWr-doped devices, the decrease of the solar conversion efficiencies at low temperatures and the associated decrease of the integrated external quantum efficiency through InGaAs could be attributed to detected traps E1QWR_D, E2QWR_D, and E3QWR_D with activation energies of 0.0037, 0.0053, and 0.041 eV, respectively.
... Moreover, noise spectroscopy is quite an effective technique when DLTS fails: levels with very small capture cross sections and levels in semi-insulators. [11][12][13] In order to conduct DLTS and low-frequency noise and G-R noise measurement in n-GaN wafer, GaN Schottky diodes were fabricated. The structure of n-GaN Schottky diode is shown in Figure 1. ...
... Capture crosssection r n was found as $10 À14 cm 2 . The density of the trap was calculated from amplitude of G-R processes in the volume of the semiconductor using following equation: 12 of the G-R process in the volume of the semiconductor, N T is the trap density, V is the sample volume, and N d is the donor concentration. The trap density was found to be N T $ 10 14 cm À3 with A ¼ 1.342 Â 10 À7 . ...
Article
Full-text available
We report on the identification of a deep level trap centre which contributes to generation-recombination noise. A n-GaN epilayer, grown by MOCVD on sapphire, was measured by deep level transient spectroscopy (DLTS) and noise spectroscopy. DLTS found 3 well documented deep levels at E[subscript c] − 0.26 eV, E[subscript c] − 0.59 eV, and E[subscript c] − 0.71 eV. The noise spectroscopy identified a generation recombination centre at E[subscript c] − 0.65 ± 0.1 eV with a recombination lifetime of 65 μs at 300 K. This level is considered to be the same as the one at E[subscript c] − 0.59 eV measured from DLTS, as they have similar trap densities and capture cross section. This result shows that some deep levels contribute to noise generation in GaN materials.
... Moreover, noise spectroscopy is quite an effective technique when DLTS fails: levels with very small capture cross sections and levels in semi-insulators. [11][12][13] In order to conduct DLTS and low-frequency noise and G-R noise measurement in n-GaN wafer, GaN Schottky diodes were fabricated. The structure of n-GaN Schottky diode is shown in Figure 1. ...
... Capture crosssection r n was found as $10 À14 cm 2 . The density of the trap was calculated from amplitude of G-R processes in the volume of the semiconductor using following equation: 12 ...
... This trap density is comparable to that of the InAs/InGaAs QDs structure grown on native GaAs substrate (in the range of 10 12 to 10 14 cm −3 ). 27,28 Finally, we want to point out that, since the dark current observed in the QD p-i-n photodiodes correlates to native defects in the materials, there is still room to further decrease the dark current with a longer carrier lifetime and lower defect density. 18,29 In addition, the carrier lifetime extracted here can be used to model the carrier dynamics process and quantum efficiency in the QD p-i-n photodiodes. ...
... LFNS has some advantages over DLTS, since the latter requires high excitation signal (reverse voltage bias), proper samples with depletion region and with small leakage currents. It is also possible with LFNS to roughly estimate the trap density, when dopant concentration and sample volume are known [8]. This problem vanishes in LFNS because current fluctuations are the source of the information about traps in contrast to DLTS, where capacitance transients are analysed. ...
Article
Full-text available
Low frequency noise spectroscopy (LFNS) is a tool, complementary to other methods, such as deep level transient spectroscopy, to characterize traps in semiconductor devices. LFNS method was described, illustrated and applied to HgCdTe high operating temperature infrared detectors. The fluctuation model which connects generation- recombination process (Lorentzians) with trap level in doped semiconductor was briefly introduced. Several trap levels were obtained for detectors with different band gaps (Eg in the range 219 meV to 320 meV at 300 K). One of the trap levels Et ≈140 meV, is common for almost all examined specimens. This level, which lies in the middle of the band gap, is the main source of generation-recombination noise observed in examined HgCdTe HOT detectors.
... LFNS is a useful method for the electrical characterization of a number of semiconductor devices [35][36][37]. However, it had never been applied to InAs/GaSb superlattice materials or devices until now. ...
Article
Full-text available
In this report, we present results of an experimental investigation of a near mid-gap trap energy level in InAs10 ML/GaSb10 ML type-II superlattices. Using thermal analysis of dark current, Fourier transform photoluminescence and low-frequency noise spectroscopy, we have examined several wafers and diodes with similar period design and the same macroscopic construction. All characterization techniques gave nearly the same value of about 140 meV independent of substrate type. Additionally, photoluminescence spectra show that the transition related to the trap centre is temperature independent. The presented methodology for thermal analysis of dark current characteristics should be useful to easily estimate the position of deep energy levels in superlattice photodiodes.
Article
This work reports on the optical properties of self-assembled InAs/GaAs quantum dots (QDs) grown by molecular beam epitaxy. A sigmoidal temperature-dependent variation of the bandgap energy of the dots is detected from the photoluminescence (PL) investigations of the studied sample. This sigmoidal dependence of the bandgap energy was accompanied by an anomalous increase in the integrated PL intensity (IPLI) over the temperature range of 10 K to 110 K. Such optical observations are related to the existence of deep localized states (LS) within the QD structure. To illustrate the presence of the LS, pump power-dependent PL measurements at different temperatures were performed. It is found that the sigmoidal dependence of the PL peak energy (PLPE) gradually converts into a monotonous variation with the gradual increase in the excitation density. These observations show a saturation behavior of a limited number of states within the bandgap of the QDs. Such optical results provide clear evidence for the existence of deep levels within the InAs dots.
Article
We report on the identification of a deep level trap centre which contributes to generation-recombination noise. A n-GaN epilayer, grown by MOCVD on sapphire, was measured by deep level transient spectroscopy (DLTS) and noise spectroscopy. DLTS found 3 well documented deep levels at Ec − 0.26 eV, Ec − 0.59 eV, and Ec − 0.71 eV. The noise spectroscopy identified a generation recombination centre at Ec − 0.65 ± 0.1 eV with a recombination lifetime of 65 μs at 300 K. This level is considered to be the same as the one at Ec − 0.59 eV measured from DLTS, as they have similar trap densities and capture cross section. This result shows that some deep levels contribute to noise generation in GaN materials.
Article
Full-text available
The relative contributions of the photon and thermal coupling mechanisms to the behavior of self-assembled InAs/GaAs quantum dot lasers are studied. A theoretical model, which takes into account a photon coupling process between the ground and first excited states of different sized dots, is proposed to fully explain the temperature dependence of the threshold current density (Jth) of both undoped and p -doped lasers. The simulation results suggest that the carrier distribution between the different energy states in a dot is modulated by the intradot thermal excitation of carriers. This process, when combined with the photon coupling mechanism, can account for the negative characteristic temperature (T0) appearing in different temperature ranges for undoped and p -doped devices. Thermal coupling, which involves thermal carrier escape and recapture among different dots, has also been studied. Below threshold, thermal coupling is found to be significant but is weakened as threshold is approached because of the decreased carrier lifetime. Near and above threshold, the photon coupling mechanism is important and can be used to model the different temperature behaviors of the lasing spectra observed experimentally for the undoped and p -doped lasers.
Article
Full-text available
A method of recharging a local level by minority carriers has been used to determine the local level parameters in n-type 6H - SiC. Measurements have been made in the temperature range 300 - 600 K. For this level the capture cross section of electrons 0268-1242/11/8/004/img6 depends exponentially on temperature: 0268-1242/11/8/004/img7, with 0268-1242/11/8/004/img8 where 0268-1242/11/8/004/img9 is the energy gap between the level position and the bottom of the conduction band. Injection through a forward-biased p - n junction has been used to introduce minority carriers (holes) into the n-region. A set of parameters: 0268-1242/11/8/004/img10, 0268-1242/11/8/004/img9, 0268-1242/11/8/004/img12 and the level concentration 0268-1242/11/8/004/img13 has been determined without any fitting procedure.
Article
Full-text available
Self-assembled strained semiconductor nanostructures have been grown on GaAs substrates to fabricate quantum dot infrared photodetectors. State-filling photoluminescence experiments have been used to probe the zero-dimensional states and revealed four atomic-like shells (s,p,d,f) with an excitonic intersublevel energy spacing which was adjusted to ∼60 meV. The lower electronic shells were populated with carriers by n doping the heterostructure, and transitions from the occupied quantum dot states to the wetting layer or to the continuum states resulted in infrared photodetection. We demonstrate broadband normal-incidence detection with a responsivity of a few hundred mA/W at a detection wavelength of ∼5 μm. © 2001 American Institute of Physics.
Article
Full-text available
The bulk low frequency noise has been investigated in the 4H silicon carbide polytype. In the temperature range of 300–550 K the noise spectral density S is proportional to f−1.5, where f is the measurement frequency. A very low noise level has been observed for the bulk noise. At 300 K, the Hooge constant α is as small as (2–4)×10−6 at f=20 Hz and α⩽5×10−7 at f=1300 Hz. At T⩾600 K, the generation-recombination noise of a local level makes the main contribution into the total low frequency noise. The electron capture cross section σ of this level depends very strongly on temperature. In the temperature range of 620–700 K, the temperature dependence of σ can be expressed as σ∼exp(−E1/kT), with an activation energy E1 as high as 2.7 eV. © 1997 American Institute of Physics.
Article
Full-text available
Capacitance–voltage characteristics have been measured at various frequencies and temperatures for structures containing a sheet of self-assembled InAs quantum dots in both n-GaAs and p-GaAs matrices. Analysis of the capacitance–voltage characteristics shows that the deposition of 1.7 ML of InAs forms quantum dots with electron levels 80 meV below the bottom of the GaAs conduction band and two heavy-hole levels at 100 and 170 meV above the top of the GaAs valence band. The carrier energy levels agree very well with the recombination energies obtained from photoluminescence spectra. © 1998 American Institute of Physics.
Article
Full-text available
We report a bias-controllable multiwavelength quantum dot infrared photodetector (QDIP). The active region of the QDIP consisted of five layers of InAs quantum dots with InGaAs cap layers. Photoresponse peaks at 5.5, 5.9, 8.9, and 10.3–10.9 μm were observed. The relative response of these peaks could be controlled through the applied bias. For 5.9 μm detection, a peak detectivity, D∗, of 5.8×109 cm Hz1/2/W at 77 K and 0.3 V was achieved. © 2002 American Institute of Physics.
Article
Self-assembled InAs quantum dots (QDs) were grown by molecular beam epitaxy (MBE) on n+-GaAs substrates, capped between 0.4 mum thick n-type GaAs layers with electron concentration of 1×1016 cm-3. The effect of rapid thermal annealing at 700 °C for 60 s on the noise properties of the structure has been investigated using Au/n-GaAs Schottky diodes as test devices. In the reference sample without containing QDs, the noise spectra show a generation-recombination (g-r) noise behavior due to a discrete energy level located about 0.51 eV below the conduction band edge. This trap is ascribed to the M4 (or EL3) trap in GaAs MBE layers, related to a chemical impurity-native defect complex. In the structure with embedded QDs, the observed g-r noise spectra are due to a midgap trap level ascribed to the EL2 trap in GaAs, which is related to the InAs QDs dissolution due to the thermal treatment.
Article
We study theoretically the behaviour of a Na or a K atom adsorbed on an Al (100) surface when an inhomogeneous electric field is applied to the system by an STM tip. The method presented in this paper overcomes some difficulties associated to the large number of atoms in the system, and avoids the use of unphysical parallel periodicities. Processes like `induced surface migration' or `atomic desorption' are discussed.
Article
We show that the temperature-dependent resonances measured in the voltage noise of a series of Si δ-doped GaAs/Al0.33Ga0.67As quantum-well samples originate from fluctuations in the free-carrier concentration of electrons as they are trapped and emitted at deep metastable defects. We identify the main defect observed as the known DX center in III-V semiconductor alloys and present evidence for the existence of previously unreported DX-like traps observed in one of the samples measured in the dark and in the others when continually illuminated with infrared radiation. Further investigation of one of these defects using the Hall effect to measure temperature-dependent changes in free-carrier concentration shows a remarkable peak in the concentration of electrons as a function of temperature when the sample is continuously illuminated. This peak can be understood by using rate equations to describe the dynamics of the possible donor states. In particular, our model indicates a finite ionization energy for the Chadi and Chang DX0-like state for these defects. We propose that the observed phenomena are closely related to the spatial distribution of impurities established during growth. In particular, one might expect a significant probability for the formation of clusters of two or three Si atoms on adjacent group-III sites, which could behave as DX-like states with considerably modified energies. © 1996 The American Physical Society.
Article
An investigation of coherently grown InP quantum dots embedded in Ga0.5In0.5P by conventional space charge spectroscopy methods is reported. Deep level transient spectroscopy (DLTS) is used to obtain quantitative information on the electron emission from the dots. The applied field is found to significantly enhance the electron emission rates as seen by shifts in the peaks towards lower temperatures with increased field. Taking the field induced barrier lowering into account, the emission energy for the one electron ground state of the dot is determined as 240±10 meV. The correlation between the measured signal and the observed electron accumulation in capacitance–voltage measurements is clearly demonstrated. Further, studies of the electron emission when the average electron population in the dots was varied show that the emission energies are modified by the coulomb charging energy. Admittance measurements as a function of temperature, bias and frequency were also performed, and the results are qualitatively explained in terms of response from the dots. These observations are consistent with the effect of the signal frequency on the measured carrier concentration profile. To complete the picture and in order to distinguish the DLTS signature of the dots from that of the deep levels in GaInP, electron traps in the barrier material were also characterized. Two main electron traps, one with an activation energy of about 950 meV and the other having an activation energy of 450 meV, were present in all the samples. © 1998 American Institute of Physics.