Vous êtes sur la page 1sur 56

Materials Science and Engineering R 94 (2015) 156

Contents lists available at ScienceDirect

Materials Science and Engineering R


journal homepage: www.elsevier.com/locate/mser

Deep traps in GaN-based structures as affecting the performance of


GaN devices
Alexander Y. Polyakov a, In-Hwan Lee b,*
a

National University of Science and Technology MISiS, Leninsky Ave. 4, Moscow 119049, Russia
School of Advanced Materials Engineering and Research Center of Advanced Materials Development, Chonbuk National University, Jeonju 561-756, Republic
of Korea
b

A R T I C L E I N F O

A B S T R A C T

Article history:
Available online 27 May 2015

New developments in theoretical studies of defects and impurities in III-Nitrides as pertinent to


compensation and recombination in these materials are discussed. New results on experimental studies
on defect states of Si, O, Mg, C, Fe in GaN, InGaN, and AlGaN are surveyed. Deep electron and hole traps
data reported for GaN and AlGaN are critically assessed. The role of deep defects in trapping in AlGaN/
GaN, InAlN/GaN structures and transistors and in degradation of transistor parameters during electrical
stress tests and after irradiation is discussed. The recent data on deep traps inuence on luminescent
efciency and degradation of characteristics of III-Nitride light emitting devices and laser diodes are
reviewed.
2015 Elsevier B.V. All rights reserved.

Guiding Editor: Steve Pearton


Keywords:
III-Nitrides
Dislocations
Deep traps
HEMT
LED
Device degradation

Contents
1.
2.

3.

4.

5.

6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Theoretical results and experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Deep states related to native defects, impurities, and their complexes in III-Nitrides (theoretical results) .
2.1.
Deep states related to dislocations (theoretical results, brief comparison with experiment) . . . . . . . . . . . .
2.2.
Experimental studies of the properties of donor and acceptor dopants in III-Nitrides . . . . . . . . . . . . . . . . . . . . . .
Si and oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Si and O related complexes in III-Nitrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Mg in III-Nitrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
Carbon in III-Nitrides (experiment) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
Carbon in GaN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1.
C in AlN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.2.
Fe doping effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.
Other deep traps experimental studies in III-Nitrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Deep traps in GaN. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Electron and hole traps in AlGaN. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Deep traps in AlGaN/GaN and in GaN-based HEMTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
AlGaN/GaN heterojunctions: the origin of two-dimensional gas and deep traps in the material . . . . . . . . .
5.1.
Trapping in transistors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Radiation effects in GaN-based HEMTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.
Deep traps in GaN-based LEDs and LDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Deep traps in GaN/InGaN and AlGaN/AlGaN LEDs and LDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.
V.2.LEDs degradation studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.

* Corresponding author.
E-mail address: ihlee@jbnu.ac.kr (I.-H. Lee).
http://dx.doi.org/10.1016/j.mser.2015.05.001
0927-796X/ 2015 Elsevier B.V. All rights reserved.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

2
3
3
7
8
8
11
11
14
14
15
16
18
18
23
25
25
31
39
44
44
48

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

7.

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Wide-bandgap III-Nitrides have grown into one of the most
important semiconductor materials systems with vast applications
in visible-UV light emitting devices (LEDs) and laser diodes (LDs),
high-power transistors and rectiers, high-frequency devices.
Deep traps in such devices can play a very important role. This is
particularly so for the III-Nitrides because these compounds are
commonly grown on lattice mismatched substrates, with a large
lattice mismatch between individual epitaxial layers. Growth is
performed at high temperatures potentially conducive to strong
impurity contamination, high concentration of point defects and
high strain caused by the difference in thermal expansion
coefcients. All this can result in high density of extended defects
and centers with deep levels. Initial deep center defects studies
were, therefore, aimed at establishing the properties of major
native defects, impurities and dopants and their complexes both
theoretically and experimentally. Also, one of the most important
tasks was to study the electronic structure of extended defects in
III-Nitrides. These studies in III-Nitrides were seriously complicated by the wide-bandgap of most nitrides necessitating the
development of new methods of deep traps spectroscopy. Indeed,
the bandgap of GaN is 3.4 eV at room temperature, while the depth
of deep traps that can be easily scanned by the standard deep levels
transient spectroscopy (DLTS) due to Lang [1] is usually 1 eV from
the conduction or valence band. Thus, the probed regions of the
bandgap even in GaN miss the most important midgap region, the
more so for wider bandgap InAlGaN ternaries and quaternaries.
This can in principle be remedied by increasing the high
temperature in the DLTS temperature scan, but the leakage
current of Schottky diodes or pn diodes in GaN-based devices is
usually relatively high even at room temperature which makes it
very difcult to carry out reliable capacitance measurements at
temperatures exceeding 500 K. Undoped III-Nitrides grow
usually preferentially n-type, so minority carriers traps in such
lms had to be studied on Schottky diodes by the DLTS version
with optical injection pulse, the so called optical DLTS (ODLTS)
[2]. The as-grown, annealed or irradiated III-Nitrides are often
semi-insulating and deep traps in them have to be detected by
using the photoinduced current transient spectroscopy (PICTS or
OTCS) which is similar to DLTS, but utilizes photocurrent
relaxations induced by optical injection pulse [36].
The III-Nitrides lms and layers are often heavily compensated
so that standard DLTS measurements on them are not reliable
[1]. In that case current version of DLTS (CDLTS) (see [7] and
references therein) is often an answer. It also helps in analysing the
spectra of deep traps in p-type nitrides because the depth of p-type
dopants in nitrides is relatively high and causes strong freeze-out
of material at temperatures below room temperature.
Admittance spectroscopy (see e.g. [8]) and capacitancevoltage
(CV) proling in the dark or under illumination (LCV) [7,9] were
found to be very useful in determining the thermal and optical
ionization energy of deep traps near the Fermi level in
compensated nitrides. These techniques also proved very useful
in analyzing the electrical properties and charge distribution in
quantum well or multi-quantum-well structures.
In order to circumvent the problem with high ionization energy
of traps in GaN and AlGaN the deep levels optical spectroscopy
(DLOS) [10] was widely adopted in nitrides studies. In this method

51
52
52

the bandgap is scanned not by sweeping the temperature, as in


DLTS, but rather by sweeping the photon energy of excitation light.
Because the dislocation density in III-Nitrides is usually quite
high methods had to be developed to detect deep traps decorating
dislocations as well as the states that belong to dislocations proper.
Here a wide application was found for the method employing DLTS
peaks amplitude measurements as a function of the injection pulse
length (due to Wosinski [11]).
Additional problems in all measurements involving light
excitation are caused by the necessity to employ very shortwavelength light sources in DLOS, ODLTS, PICTS, photoluminescense (PL) spectroscopy. The results of the initial round of these
theoretical and experimental studies have been summarized in
several reviews [1218]. Theory pointed to the dominant role of
gallium vacancy (VGa) acceptors in n-type nitrides and nitrogen
vacancy donors (VN) in p-type nitrides. Experimental studies were
able to map out the major electron and hole traps in n-GaN and
partly in p-GaN. Identication of the possible origin of the traps,
when it was done, understandably relied very heavily on the
results of radiation studies which was reected in the general tenor
of the review articles. As a pleasant surprise it was found that the
densities of deep traps in nitrides are not as high as could be
anticipated and the dislocations are far less detrimental to the
devices performance than for other III-V materials. The contribution of ShockleyReadHall (SRC) recombination in nitride light
emitters was not found to be the major factor. At the same time,
radiation experiments showed that the radiation tolerance of III-N
devices is one-two orders of magnitude higher than for Si or GaAsbased devices and does not pose an immediate problem in device
applications. Hence, the somewhat decreased interest to defects
with deep centers in nitrides in recent years.
However, lately the new understanding of the role of deep traps
in III-Nitride devices starts to emerge. It is based both on the results
of new theoretical calculations and on thorough investigations of
the role of deep traps in performance of LEDs, LDs, HEMTs and in
their degradation under electrical stress. For example, theoretical
work has re-assessed the role of C in III-Nitrides showing it to be a
deep trap impurity with the midgap acceptor level and a deep
donor level. For the main shallow acceptor, Mg, theory shows that
it, in fact, has all the characteristics of deep level defects. These
ndings have potentially profound consequences for interpretation of compensation mechanisms in III-Nitrides (very important
for optimizing growth of semi-insulating buffers in HEMTs). They
are also important for attribution of defect related PL bands in IIINitrides which is of great interest in understanding the behavior of
UV-LEDs and their degradation.
Recent experiments point to a serious role of deep trap defects
in determining the quantum efciency and the threshold current of
LEDs and LDs, trapping in HEMTs and the way these devices
deteriorate when operated under high driving current conditions
typical of their practical use. It seems that traps decorating
dislocations play a particularly prominent part in these effects. A
new understanding is also accumulating on possible links between
the changes induced by irradiation and the degradation of devices
during electrical stress testing. It seems that these efforts have to
be systematized in order to coordinate further studies in that
direction. This is the major aim of the present review.
The structure of the paper is as follows. In Section 2, we briey
describe the previously obtained results of theoretical calculations

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

for the states of defects and impurities and different types of


dislocations in III-Nitrides to provide an introduction to the eld
for uninitiated reader, but the emphasis is on new theoretical
results discussing the role of C and Mg in III-Nitrides and the
decoration of dislocations by Ga, In, Al. Section 3 summarizes the
known results on behavior of the main technologically important
impurities in III-Nitrides: Si and O (shallow donor dopants, main
donor contaminants, metastable DX-like centers), Mg (the main
acceptor in GaN and ternary solid solutions of various composition), C (main compensating impurity used to obtain semiinsulating (SI) buffers in HEMTs, very important center for the
performance of UV-LEDs), and Fe (the main deep center used for
donor conductivity compensation in SI-GaN buffers of HEMTs). In
Section 4, we systematize the known and recent data on the
spectra of deep electron and hole traps in GaN and AlGaN. Section 5
describes the behavior of deep traps in HEMT structures and
HEMTs and the traps roles in degradation and after irradiation.
Section 6 treats the role of deep traps in luminescence efciency
and degradation of III-Nitride LEDs and LDs.
2. Theoretical results and experiment
2.1. Deep states related to native defects, impurities, and their
complexes in III-Nitrides (theoretical results)
The formation energy and energy level positions for major
native defects in III-Nitrides were analyzed in a number of papers
(see Refs. [12,1926] to name a few). Excellent summary of the
theoretical work performed on this front up to middle 2000s has
been done in Refs. [12,27]. It appears that, when one takes into
account the formation energies of various native defects in IIINitrides, only Ga vacancies (VGa) in n-type materials and N
vacancies (VN) in p-type materials should be encountered in
signicant concentrations in as-grown lms and crystals, other
defects, cation and anion interstitials, antisite defects, VGa in p-type
and VN in n-type materials having high formation energies [12,20].
Figs. 1 and 2 taken from Ref. [12] show the Fermi energy
dependence of the formation energy and the charge states and
respective energy levels for the main simple native defects in GaN.

Fig. 1. Formation energies as a function of Fermi level for native point defects in
GaN. Ga-rich conditions are assumed. The zero of Fermi level corresponds to the top
of the valence band. Only segments corresponding to the lowest-energy charge
states are shown. The slope of these segments indicates the charge state. Kinks in
the curves indicate transitions between different charge states.
(After Ref. [12], Fig. 5) Copyright American Institute of Physics 2004.

Fig. 2. Thermodynamic transition levels for defects in GaN, determined from the
formation energies displayed in Fig. 1.
(After Ref. [12], Fig. 6) Copyright American Institute of Physics 2004.

VGa are triple acceptors lled up to the VGa3 level in n-GaN, VN are
donors giving rise to shallow singly positively charged states in nGaN and deep triply positively charged states VN3+ with a level in
the lower half of the bandgap in p-GaN (the doubly charged state is
metastable and should not be observed in equilibrium conditions
[12,20,21]). Ni and Gai in Fig. 1 are, respectively, nitrogen
interstitials (deep acceptors in n-GaN, deep donors in p-GaN, see
Figs. 1 and 2) and Ga interstitials (deep donors). NGa and GaN are
the nitrogen and gallium antisite defects. The zero of Fermi energy
in Fig. 1 corresponds to the valence band maximum, positive slope
of the energy dependence is for donor states, negative slope is for
acceptor states, the points where the slope changes are the
energies at which the defects change their charge state [12]. Naturally, under strongly nonequilibrium conditions (e.g. irradiation
with high energy particles or device operation under very high
power), defects other than vacancies can be observed, but antisite
defects do not seem to be playing any important role even under
these conditions (we will discuss that later).
Switching from GaN to InN or AlN does not change the general
picture qualitatively (see [2227]), although it should be noted
that the calculations for these latter binaries were mostly
performed for the zinc-blend rather than wurtzite polytype. For
InN, an additional uncertainty is added by the fact the bandgap
energy calculations predict the bandgap that is much lower than
the experimentally observed value of about 0.8 eV and special
corrections have to be introduced to circumvent this problem
[12,25]. Fig. 3 taken from Ref. [25] illustrates the Fermi energy
dependence of the defects formation energies in InN. It can be seen
that the lowest energy defects are the gallium and nitrogen
vacancies in respectively n- and p-InN, the formation energy of
nitrogen antisites NIn and nitrogen interstitials Ni is always very
high. However, in contrast to GaN, the formation energies of In
interstitials Ini and In antisites InN in p-InN is low and comparable
to the formation energy of the VN3+ donors.
For AlN, again, the lowest energy defects turn out to be the
triply negatively charged Al vacancies in n-type material and triply
positively charged nitrogen vacancies in p-type material [26]. However, the formation energy of Al interstitials, donors in p-AlN, was
found to be comparable to the formation energy of the main
compensating defect in p-type III-Nitrides, the triply positively
charged VN (in fact, the formation energy for Ali is even lower).
Fig. 4 taken from Ref. [26] depicts the energy positions for different
defects as calculated for zinc-blend AlN. It should be noted here
that, for wurtzite AlN, because of the different coordination of
defects and the larger bandgap than for the zinc-blend polytype,
the general picture would be seriously different [12,26]. For
example, the VN+ state that is a resonance in zinc-blend AlN
produces a deep level in wurtzite AlN, the energy level of the VN3+

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 3. The Fermi energy dependence of the simple native defects formation energy
in InN.
(After Ref. [25], Fig. 1) Copyright the American Physical Society, 2000.

donor is about Ev+1 eV (compared to Ev+0.2 eV in GaN) rather


than close to the valence band maximum as for the zinc-blend AlN,
the formation energy of the Ali donors in the wurtzite polytype is
considerably higher than in the zinc-blend polytype [26] (it is
suspected that similar changes in Ini formation energies as a
function of polytype would be observed for InN [12]).
Theoretical calculations for simple substitutional donors, Si and
Ge on the Ga site, oxygen on the nitrogen site in GaN, predict a high
solubility, low ionization energy and the lack of strong lattice
relaxation effects resulting in metastabilities [12,28,29]. For AlN,
the results differ depending on the type of donors. For O, it is
unanimously expected that the center will experience the DX-like
transformation turning oxygen from a simple donor as in GaN into
a deep metastable center [12,2830]. For Si, the predictions of
different groups differ. Van de Walle [29] concludes that Si should
not form a deep DX-like state in AlN, whereas the authors of Ref.
[30] expect Si to be a metastable DX-like defect. For Ge Park et al.
[28] predict the center in AlN to be a simple donor, while
Boguslawski and Bernholz [31] expect to observe DX-like behavior.
Hydrogen is an extremely important impurity in III-Nitrides
because it can be easily introduced in high concentrations during
growth and processing and form stable complexes with dopants
and defects, thus radically affecting their electrical properties (see
e.g. a recent review in Ref. [32] and references therein, we will be
discussing these matters later in the paper). Theoretical studies of
hydrogen electronic states in GaN have been performed in a
number of papers [3338], a very good summary of the results can
be found in Ref. [12]. In Refs. [39,40] the idea of hydrogen levels
band alignment in respect to the level of vacuum was used to
predict the behavior of hydrogen in a variety of semiconductor
materials including GaN, AlN, InN. The actual detailed calculations
for InN and AlN and InN were performed in Refs. [34,40]. Briey,
the results are as follows. In GaN hydrogen is an amphoteric
impurity producing a donor level H+ in p-GaN and an acceptor level

Fig. 4. The calculated energy levels of simple native defects in zinc-blend AlN.
(After Ref. [26], Fig. 2) Copyright The American Physical Society, 2002.

Fig. 5. The dependence of the formation energy of hydrogen donors and hydrogen
acceptors in GaN on the Fermi level position.
(After Ref. [12], Fig. 15) Copyright American Institute of Physics 2004.

H in n-GaN. Fig. 5 taken from Ref. [12] shows the dependence of


the hydrogen formation energy in different charge states on the
Fermi level position in GaN. It can be seen that the formation
energy is expected to be very much lower and hence the hydrogen
concentration to be very much higher in p-GaN than in n-GaN. The
gure also illustrates the point that neutral hydrogen is never
thermodynamically stable in GaN. Mind that for GaN both states of
hydrogen can be observed depending on the Fermi level position.
The same is true for AlN [40]. In contrast, for InN the cross-over
point between the Fermi energy dependence of the formation
energy for the H+ and H states lies well above the conduction band
edge, so that hydrogen is always a shallow donor [34,39,40].
As for the acceptors, the main shallow acceptor in III-Nitrides is
Mg which is well accounted for by theory (see an excellent review
in Ref. [12]). The Mg acceptor level is located close to Ev+(0.2
0.3) eV in GaN and InN, but is much deeper, near Ev+0.5 eV in AlN
[12]. Recent detailed studies of the electronic structure of Mg
acceptors in GaN, AlN, and InN [39] revealed that, in GaN, the Mg
acceptors in fact behave as deep acceptors with a highly localized
hole, in contrast to the true shallow acceptors in which the hole is
delocalized and its behavior is described by the effective mass
theory. The fact that the Mg acceptor level near Ev+0.26 eV is
relatively shallow is believed to be simply a coincidence. This is
even more the case for AlN with Mg acceptor level located close to
Ev+0.5 eV. However, in InN the behavior of Mg acceptors
corresponds to that expected for a true shallow acceptor near
Ev+0.19 eV [39].
Theoretical studies for alternative acceptors, Be, Ca, Zn, C, are
reviewed in Ref. [12]. Neither of these acceptors has been
consistently and successfully used as a shallow p-type dopant in
III-Nitride devices, for which reason we shall not discuss their
behavior in detail in this paper. The exception is carbon considered
here owing to its importance for obtaining high resistivity in GaN
HEMT buffers and for the formation of deep absorption bands
handicapping the performance of UV LEDs with compositions close
to AlN (we will discuss these phenomena later). Theoretical studies
of the electronic properties of C in GaN have been published in Refs.
[12,23,26,31,41,42]. In the early publications [12,23,26,31,41] it
was argued that C on N site, CN, should be a shallow acceptor
located near Ev+0.2 eV, C on Ga site should be a shallow donor,
interstitial C, Ci, should be a deep midgap center behaving as a deep
compensating acceptor in n-GaN and as a deep compensating
donor in p-GaN. The situation was reassessed in Ref. [42] where the
authors show that CN in GaN is in fact a deep acceptor with a level
near Ev+0.9 eV. This acceptor is held responsible for compensation

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 6. The Fermi level dependence of the formation energy of the three main C
species in GaN under Ga-rich (a) and N-rich (b) growth conditions.
(After Ref. [43], Fig. 1) Copyright American Physical Society 2014.

and high resistivity in C-doped n-GaN. Formation energy


considerations made the authors of Ref. [42] discard the possibility
of self-compensation in GaN via the formation of CN and CGa or CN
and Ci in close concentration. Further studies were performed in
Ref. [43] to include C doping effects in AlN and InN. Fig. 6 taken
from Ref. [43] illustrates the Fermi level dependence of the
formation energy of the three main C species in GaN for Ga-rich
and N-rich growth conditions. Mind that, in n-GaN, the Ev+0.9 eV
acceptors due to CN should be absolutely dominant, the formation
energy of the shallow C donors and the deep Ci centers being too
high. The new additional feature for the CN states, as compared to
Ref. [42], is the appearance of a donor level near Ev+0.35 eV
(positive slope in respective line in Fig. 6) in p-type GaN. These
centers should give rise to metastable behavior depending on the
steady-state Fermi level position and on illumination intensity.
Shallow CGa donors and deep Ci donors dominate in p-type GaN
and this dominance becomes very pronounced under N-rich
conditions characteristic of MOCVD and HVPE growth techniques.
Thus, in such materials compensation of the major acceptors is
provided by the donor species of C. Under Ga-rich conditions
typical of MBE growth all three species, the CN donors near
Ev+0.35 eV, the shallow CGa donors, and the Ci donors contribute to
compensation in p-GaN. It is interesting also that one expects to
observe deep states at Ev+2 eV associated with Ci in n-GaN
although they should not be important in equilibrium. We will see
below that centers with similar energies are quite prominent in
deep traps spectra of some of the C doped n-GaN lms.
The situation in AlN is somewhat different from that in GaN. The
CN states are still expected to be the dominant compensating
acceptors in n-AlN, but the respective acceptor level is much
deeper than in GaN, close to Ev+2 eV, the donor level of CN is also
deeper than in GaN, close to Ev+1 eV. CAl is no longer a shallow
donor. It forms an acceptor level at Ec1.68 eV and a donor level at
Ec1.82 eV. The latter donors should be the major compensating
centers in p-AlN grown under N-rich conditions, but not for the Garich conditions. Interstitial Ci is, as in GaN, either a doubly charged
donor or a doubly charged acceptor, but its formation energy is
always much higher than for other C species so that Ci should not
dominate in electrical properties in equilibrium conditions [43].
In InN, the depth of the CN acceptors is close to Ev+0.59 eV, there
is no donor state associated with CN, the CN acceptors are expected
to be dominant compensating centers only in heavily degenerate
n-type InN with a strong BursteinMoss shift (see e.g. Ref. [44]).

Fig. 7. The band alignment of the CN acceptor (solid lines) and donor (dashed lines)
levels in GaN, AlN, and InN.
(After Ref. [43], Fig. 8) Copyright American Physical Society 2014.

Both, CIn and Ci form shallow donor states, and the shallow donors
due to doubly positively charged CIn 2+ donors are predicted to
dominate in lightly doped InN [43]. The changes in the energy
levels positions of C states in GaN, AlN, InN support the idea that
the levels are aligned in respect to the vacuum level, as illustrated
for the CN acceptor and donor states in Fig. 7 taken from Ref. [43].
As we have seen above, native defects in III-Nitrides give rise to
a large variety of donor and acceptor states and thus can form
complexes between oppositely charged centers and with impurity
atoms. By far the most important among such complexes are
acceptor Mg and donor hydrogen pairs because their formation
handicaps efcient p-type doping of nitrides. Kinetics of hydrogen
diffusion in different charge states, microscopic structure and the
binding energy of MgH complexes (believed to be neutral) were
analyzed in Refs. [25,26,37,38]. These papers also treat variations
in H concentration observed for varying Mg concentration, the
hydrogen outdiffusion and the kinetics of hydrogen release upon
annealing. This theoretical analysis showed that the binding
energy of MgH complexes in p-GaN is close to 0.71 eV, the
presence of hydrogen suppresses the formation of compensating
VN3+ defects and increases the Mg solubility (see also a review in
Ref. [12]). Some additional features of the process were discovered
in a recent analysis of the phenomena in Ref. [39]. Fig. 8(a) taken
from Ref. [39] illustrates the major differences with the earlier
picture. It can be seen from the Fermi level dependence of the
formation energies of the Mg acceptors MgGa, isolated interstitial
hydrogen Hi, and MgH complexes that, in addition to neutral Mg
H complexes (the authors obtained the binding energy of these
complexes as 1.02 eV), one should observe in p-type material a
donor state near Ev+0.13 eV. For AlN, the authors of Ref. [39]
predict the donor state of the MgH complex to be deeper, close to
Ev+0.43 eV. All that has serious implications for possible optical
transitions in the material as will be discussed below. Mg acceptors
complexing with other donors, nitrogen vacancies in GaN, was
analyzed in Ref. [45]. It was shown that such complexes can create
a donor state near Ev+0.9 eV, more prominent under Ga-rich
growth conditions.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 8. (a) The Fermi level dependence of the formation energy of Mg acceptors, interstitial hydrogen and hydrogenMg complexes in GaN, (b) conguration-coordinate
diagram for optical transitions via Mg acceptors in GaN.
(After Ref. [39], Fig. 2) Copyright American Physical Society 2012.

Interstitial hydrogen donors can also form complexes with


other acceptors in GaN, such as Be and CN. The binding energy of
the complexes is close to 1 eV [9,22]. For CNH complexes, it was
found in Ref. [39] that a donor state near Ev+0.25 eV can be formed
in p-GaN. In Ref. [46] it has been proposed that complexes of
substitutional CN in GaN with another prominent donor, oxygen,
can form an acceptor state with level close to Ev+1 eV, although it
has been argued in Ref. [43] that such complexes would have a very
low binding energy.
Another very important example of donoracceptor complexes
are complexes of Ga vacancy acceptors with O or Si shallow donors
in GaN. Such complexes were theoretically studied in Refs. [12,21
24]. It is predicted that complexes with O are more thermally
stable than with Si and that respective defects form acceptor states
close to the acceptor states of isolated Ga vacancies, but instead of
triply charged acceptors doubly charged acceptors are formed.
Similar behavior for VAlO complexes was predicted in Ref. [22] for
AlN.
Predicting the type of optical absorption (OA) and photoluminescence (PL) bands associated with certain defect states is an
important part of theoretical studies of defects. This work was
done in a number of papers. The most telling examples are the
analysis of OA and PL bands in Mg doped, C doped, O doped IIINitrides. For Mg acceptors in GaN theoretically calculated
conguration-coordinate diagram shown in Fig. 8(b) taken from
Ref. [43] predicts the appearance of a broad blue band near 2.7 eV
that is often observed experimentally, particularly for annealed
MOCVD grown material and MBE grown material in which the
concentration of the MgH complexes is low. On the other hand,
the MgH complexes producing a relatively shallow donor state
near Ev+0.13 eV show weak lattice relaxation effects. Corresponding PL line is predicted to be narrow and peak at 3.32 eV [43]. For
isolated Mg acceptors in AlN, the calculated congurationcoordinate diagrams predict the existence of a wide violet band
peaked at 4.77 eV and analogous in nature to the blue band in pGaN [43]. In addition, recombination with the donor state of Mg
acceptors in AlN should produce a broad PL band at 5.43 eV, whilst,
for recombination involving MgH complexes in AlN, the presence
of a deep donor state near Ev+0.43 eV resulting in PL band peaked
at 5.13 eV is expected. In contrast, for InN, the Mg acceptor is a true
shallow acceptor producing no strong lattice relaxation, while the
MgH complexes are neutral [43]. For Mg acceptor complexes with
nitrogen vacancy donors in GaN, theory [45] predicts the existence
of a broad red PL band peaked near 1.8 eV. These ndings are
important when GaN, AlN, AlGaN, and InGaN are doped with Mg
(to be discussed below).

Another important example is the broad defect yellow PL band


in GaN and similar bands in AlN and AlGaN (see e.g. a discussion in
Ref. [27]). This example shows the extreme difculty of the
situation. The results offered in various theoretical publications are
often contradictory. In Refs. [2123] the yellow luminescence band
is attributed to recombination involving VGaO acceptor complexes with the doubly negatively charged level near Ev+1 eV.
However, in Ref. [42,43] respective optical transition is attributed
to recombination on the Ev+0.9 eV CN acceptors. The calculated
conguration-coordinate diagrams place the PL peak of this
transition at 2.14 eV, while the existence of the donor state of
CN gives rise to a broad PL band near 2.7 eV [43]. For CNH
complexes a somewhat shifted PL band peaked at 2.76 eV is
predicted [43]. At the same time, an alternative attribution
involving transitions via CNO acceptor state near Ev+1 eV was
proposed for the yellow luminescence band in Ref. [46].
For AlN, a strong absorption band at 4.83 eV and broad PL band
centered at 3.69 eV are predicted for transitions involving CN
acceptor states [43]. But, as discussed above, there also exists a CN+
donor state close to the valence band maximum. The presence of
this state should give rise to additional absorption band at 5.66 eV
and PL band at 4.5 eV [43].
Finally, it should be discussed how the properties of defect
centers in ternary III-Nitride solutions are expected to change with
composition. We have seen, for example, that the behavior of H and
C radically changes when going from GaN to InN. For hydrogen, one
expects to see hydrogen donors or hydrogen acceptors in,
respectively, p-type or n-type materials, whereas in InN hydrogen
is always a shallow donor contributing to compensation in p-InN
and to n-conductivity in n-InN (see the discussion in Ref. [12]).
Therefore, in InxGa1xN, there should exist a boundary between the
GaN-like solutions and the InN-like solutions. It has been shown
that the hydrogen levels alignment in respect to the vacuum level
and the existence of the so called electrical neutrality level are the
concepts reasonably describing the actual situation [40]. According
to authors of Ref. [47] the electron neutrality level in all III-Nitrides
should be located 4.9 eV below the vacuum level. Based on this
concept the authors of Ref. [47] predict the boundary between the
GaN-like and InN-like states in InGaN to correspond to the In
composition close to xIn = 0.34. Similarly, for C doping, there exists
a qualitative difference between GaN and InN in that the CNacceptors dominate in the former while the Ci donors are dominant
in the latter [43]. The same goes for Mg acceptors in GaN and InN
that are deep centers with acceptor and donor states near the
valence band in GaN and true shallow acceptors in InN [39]. On
the strength of the same arguments the boundary In composition

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

should be close to xIn = 0.34 in both cases. The concept also seems
to work for predicting the defects energy evolution in AlxGa1xN.
For these ternaries, there is also a question of the border Al
composition corresponding to switching between the simple
shallow donors and DX-like behavior for oxygen, Si, Ge. Calculations performed in Refs. [2831] place this transition for O at Al
composition close to xAl = (0.30.4), for Si at xAl > 0.6, for Ge at
xAl > 0.3, although there is considerable discrepancy between the
results of different calculations (for example, Van de Walle [29]
does not expect Si to form DX-like centers for any aluminum
composition; see also an extensive discussion in Ref. [12]).
In the mind of non-theorists, to whom the authors of this review
belong, there always is, of course, the looming question of how
accurate the predictions of the theory regarding the formation
energies and the energy levels of defects are. The estimate of the
calculated energies accuracy is discussed in much detail in Ref.
[12]. It is claimed that the energies can be calculated with the
accuracy of 0.1 eV, although this reects the conversion of the
results with respect to increasing the size of the atoms cluster used.
The real accuracy is hard to estimate reliably and here is where
the experiment should come in. The serious problem, of course, is the
tendency of modern theoretical approaches to underestimate
the bandgaps of the host materials. However, supposedly the relative
position of the levels is determined more accurately than the absolute
energy and also new computational methods allow to considerably
improve the accuracy of the bandgaps calculations [12].
2.2. Deep states related to dislocations (theoretical results, brief
comparison with experiment)
The results of different theoretical calculations for electronic
states of dislocations in GaN and other III-Nitrides are somewhat
contradictory. The early work on electronic properties of screw and
edge threading dislocations in GaN [48] suggested that full core
screw dislocations should produce strong lattice distortions and
form a band of deep states near Ev+(0.91.6) eV and relatively
shallow states near Ec0.2 eV. Open core screw dislocations
formed by {10-10}facets were expected to form only shallow states
which was also a prediction for full-core edge dislocations
[48]. When comparing the formation energies of the full-core
versus open-core threading dislocations it was concluded that
open-core screw dislocations with a small diameter of around 7.9 A
could be energetically favorable, whilst for higher diameters opencore dislocations should not be observed. However, in later work of
the same group it was argued that the strain eld related to
threading dislocations is strong enough to create attractive force
for impurity segregation, oxygen being the most likely technologically important impurity to be segregated [49,50]. Complexing of
oxygen donors with triply charged gallium vacancy VGa acceptors
could then form doubly charged acceptors (VGaO)2 which in nGaN would be lled and form the acceptor transition level near
Ev+1 eV. Complexes with two oxygen atoms would produce singly
ionized acceptors, while complexes with three O atoms would be
electrically neutral and could stabilize the surface of the open-core
screw dislocations and make it possible for the large diameter
microtubes associated with open-core screw dislocations to be
formed as observed in some experiments (we shall discuss the
matter below) [50]. In contrast, for edge dislocations, the full-core
dislocations should be dominant [48,50]. Thus, in this approach the
dislocations per se are expected to produce deep gap states only for
the full-core screw dislocations, while, for open-core screw
dislocations and full-core edge dislocations, deep states near
Ev+1 eV would appear only as a result of dislocations decoration by
VGaO complexes. This approach was taken when analyzing the
effect of dislocations in n-GaN on electron mobility and produced a
reasonable agreement with the experiment [51,52].

In a later work Northrup [53] has argued that, under the Ga-rich
growth conditions characteristic of MBE growth of GaN lms,
screw dislocations with Ga-lled core should be the dominant
defects. The Ga-lled dislocation produces metallic-like density of
states over the entire bandgap, the Fermi level in the dislocation
core is pinned near Ev+1.4 eV. Thus, in doped materials dislocations
give rise to a strong band bending and the appearance of a space
charge region around them. Their contribution to electrical
properties is via the excessive leakage current through the
metallic-like core, they also are expected to serve as efcient
nonradiative recombination centers in GaN.
The possible role of threading edge dislocations has been
revisited in a number of papers. In Ref. [54] the atomic and
electronic structures of the threading edge dislocations of GaN was
studied using self-consistent-charge density-functional tightbinding approaches. Full-core, open-core, Ga-vacancy-decorated,
and N-vacancy-decorated edge dislocations were assumed to be
fully relaxed in the total-energy calculations scheme. The Gavacancy dislocation was found to be the most stable in a wide
range of Ga chemical potentials, whereas full-core and open-core
dislocations were more stable than others in the Ga-rich region.
The dangling bonds at Ga atoms were mostly contributing to the
deep-gap states, whereas those at N atoms contributed to the
formation of the valence-band tails. All the edge dislocations could
act as deep trap centers, except the Ga-vacancy dislocation. The
deep levels scheme proposed in the paper is reproduced in Fig. 9. It
can be seen that the full-core dislocations produce a whole
spectrum of defect states between Ec0.4 eV and Ec1.4 eV, opencore edge dislocations give rise to deep states near Ec0.4 eV and
Ec2 eV, dislocations decorated with Ga vacancies are associated
with relatively shallow states near Ec0.2 eV and Ev+0.2 eV,
dislocations decorated with N vacancies produce a whole range
of deep states from Ev+0.6 eV to Ec0.3 eV.
In another work on electronic structure of edge dislocations
published by Wright and Grossner [55] the authors predict that a
structure having Ga vacancies in the dislocation core should be
most stable for n-type samples grown under N-rich conditions. For
p-type and Ga-rich conditions the most stable structure is
expected to be the structure with nitrogen vacancies in the core.
Thus the electronic properties of the edge dislocations are expected
to be similar to the electronic properties of respective vacancies
decorating the core of the dislocation.
In one of the recent theoretical studies of electronic properties
of the edge dislocations in GaN [56] it is argued that such
dislocations can produce the deep acceptor states in the lower half
of the bandgap and the donor states band located about 0.5 eV
higher in the bandgap.
Interestingly, Lymperakis et al. [57] have shown that, even
under the conditions when the atomic reconstruction in the
threading dislocations core results in the formation of no broken
bonds, the high strain eld of the dislocation in combination with
the small lattice parameter of GaN can induce a metal-like
structure caused by the formation of GaGa bonds similar to bonds
in metallic Ga. The electronic structure of such dislocations would
be similar to the Ga-lled core screw dislocations described by
Northrup [53] with the difference that these strain-induced
metallic states will be insensitive to the actual growth stoichiometry and can be observed in crystals grown using techniques in
which the preferred growth conditions are, in fact, N-rich, such as
MOCVD or HVPE.
For other III-Nitrides theoretical studies of electronic properties
of dislocations were less extensive. In Ref. [58] it was suggested
that, for screw dislocations, the total formation energy arguments
favor Al segregation in the cores of dislocations in AlN and In
segregation in the cores of dislocations in InGaN and InAlN. For AlN
it can also be expected that the dislocations could be decorated by

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 9. The schematic diagram of energy levels in the bandgap of GaN for various types of edge dislocations, full-core, open-core, Ga-vacancy terminated, N-vacancy
terminated, lled circles show the occupied states in the two latter cases.
(After Ref. [54], Fig. 7) Copyright the American Physical Society, 2000.

vacancy defects, Al vacancies in n-type material and N vacancies in


p-type material (see e.g. a discussion in Ref. [12]). Hence, the
electronic properties of dislocations in p-AlN and n-AlN would be
determined, in this case, by either VN+ or VN3+ deep donors located
about 2 eV below the conduction band edge or 1 eV above the
valence band edge or by VAl3 acceptors with the level 2 eV above
the valence band edge (see Ref. [26]). Decoration of dislocations by
VGaO complexes has been directly observed in positron annihilation experiments (see e.g. Ref. [59]).
All in all, this brief discussion shows that the picture emerging
from theoretical studies is rather contradictory. Comparison with
existing experimental data to be discussed below suggests that the
deep traps that can be associated with dislocations are often
similar in energy levels position to the deep traps due to point
defects in bulk lms, suggesting that point defects decoration of
dislocations is one of the main factors in forming the electronic
structure of dislocations. As will be seen, many deep traps in GaN
and in GaN/InGaN structures show correlation with the dislocation
density. Their DLTS peaks amplitude demonstrates the logarithmic
dependence on the injection pulse length expected for correlated
charge capture by defects on dislocation line [11]. At the same
time, particularly when analysing the leakage current attributable
to dislocations in HEMTs and LEDs, one observes strong indications
of the presence of Ga, Al or In-core lled dislocations in GaN, InGaN,
AlGaN, and InAlN [6062]. The main effects here are the excessive
leakage via the metal-like dislocation core and the local changes of
the barrier height in Schottky diodes and pn junctions as caused
by the existence of such channels. Theoretical studies performed so
far do not answer the question of whether or not localized states
should be created by In segregation. However, theory suggests that
the dislocations in InGaN, InAlN should be surrounded by the Inrich regions [58]. Hence, in the case of InGaN, one would expect the
appearance of quantum-wire-like features related to In-rich lowerbandgap regions adjacent to screw dislocations. These features
could manifest themselves in PL spectra and in DLTS-like
techniques. For InAlN and for Al segregation in AlGaN the situation
is less straightforward.
Direct probing of dislocations by atomic force microscopy and
conductive atomic force microscopy show that threading edge
dislocations in GaN are commonly surrounded by negative space
charge regions and can play the role of non-radiative recombination centers while the screw dislocations mostly contribute to
leakage (see e.g. Ref. [63]). Particularly detrimental in enhanced
leakage are the open-core screw dislocations [63], especially if

they are terminated by hexagonal pits framed by the {10-11}


planes, so called V-defects [64]. Similarly work nanopipes in
MOCVD and HVPE GaN [65] or micropipes in HVPE GaN
[66,67]. The reason for excessive leakage is always due to
impurity segregation at the sidewalls of such defects [67,68]. Excessive leakage through such dislocation sites can precipitate
local overheating and consequent In segregation in LEDs (see e.g.
Ref. [69]).
3. Experimental studies of the properties of donor and acceptor
dopants in III-Nitrides
3.1. Si and oxygen
As pointed out by theoretical calculations (see above) Si on Ga
site and O on N site should be shallow donors in GaN with low
formation energy and high solubility. Experimentally it was indeed
found that both impurities can be easily incorporated into GaN up
to concentrations exceeding 1019 cm3 without causing any
serious structural degradation or the formation of impurity
precipitates (see e.g. Ref. [70]). The Si and O donor levels were
placed by the authors of Ref. [70] near Ec0.014 eV and
Ec0.028 eV, respectively. Somewhat different values of activation
energies were reported in Ref. [71]: 30 meV for Si, 33 meV for O.
Under normal pressure these are well-behaved hydrogen-like
donors causing only weak temperature dependence of electron
concentration. However, for oxygen, it was shown by Raman
spectra of free electrons measurements under hydrostatic pressure
that at a certain threshold pressure a strong freeze-out of electrons
occurs [72]. This was explained by assuming that oxygen donors in
GaN under high pressure are moved out of normal substitutional
position toward the interstitial site so that the total energy
decreases upon capture of an additional electron and the formation
of deep DX-like acceptors similar to GaAs and AlGaAs according to
reaction [28,29]:
2d ! d DX ;

(1)
+

where d8 are neutral shallow donors, d are ionized donors and


DX are deep DX-like acceptors. This happens because under
applied pressure the bandgap increases and the conduction band
edge moves upwards so that it becomes energetically favorable for
the DX state to emerge in the bandgap. Because similar changes
occur when the AlGaN ternary solution is formed it was argued
that in AlGaN of a certain threshold composition oxygen will

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

350
300

Ea(Tanyasu)
Ea(Polyakov)
Ea(Nakarmi)
Ea(Zeisel)

Ea (meV)

250
200
150
100
50
0

Fig. 10. Composition dependence of the oxygen DX levels depth in AlGaN.
(After Ref. [73], Fig. 2) Copyright the American Physical Society, 1998.

undergo the same transition from the simple hydrogen-like


substitutional donor into deep DX-like acceptor. From comparing
the pressure and composition dependences of the bandgap in the
AlGaN system it was predicted that, in AlGaN, oxygen will become
a deep DX-like acceptor at the Al mole fraction of x = 0.4 [72]. This
prediction was slightly corrected in Refs. [29,73] to take into
account the different rate of change with pressure of the
conduction band edge and the valence band edge in GaN, so that
the Al mole fraction corresponding to this shallow-deep DX-like
transition was placed near Al mole fraction of x = 0.27 (see Fig. 10
from Ref. [73]). This correlated very reasonably with the observed
emergence of the deep states in the bandgap of undoped AlGaN
ternaries reported in Refs. [74,75]. Detailed experimental and
theoretical studies performed in Ref. [73] for O-doped and
undoped n-AlGaN lms with x = 0.39 showed that, in these lms,
the centers providing electrons indeed become relatively deep and
grow deeper with increased Al mole fraction (Fig. 10 from Ref.
[73]), and that such samples are characterized by a strong
persistent photoconductivity (PPC) at low temperatures coming
from the existence of a relatively high barrier for capture of
electron by the center that has been photoionized. Such a barrier
appears because the lattice positions for negatively charged and
positively charged centers considerably differ so that the electron
capture/emission are accompanied by sizable lattice relaxation.
The conguration-coordinate diagram (CCD) was determined from
experimental dependence of the measured PPC spectra and the
measured temperature dependence of dark electron concentration
and PPC concentration. For AlGaN with x = 0.39 the optical
threshold of photoionization was found to be close to 1.3 eV,
the thermal ionization energy of DX centers was close to 0.15 eV,
and the barrier for capture of electrons was close to 0.40.5 eV
[73]. This diagram was in good agreement with the theoretically
calculated CCD for AlN [29]. As the Al composition increases the
centers become deeper and oxygen can no longer serve as an
efcient source of conduction electrons, so that undoped high-Alcomposition AlGaN lms and undoped AlN lms and crystals in
which oxygen is an important contaminant are usually highly
resistive.
For Si hydrostatic pressure experiments with GaN in Ref. [72]
did not show the shallow donor-deep DX acceptor conversion
suggesting that, if such a conversion occurs in AlGaN, the threshold
Al mole fraction should be higher than x  0.6. Indeed, experiments with silicon doping of AlGaN lms of various compositions
revealed no serious problems in attaining high electron concentration for compositions x < 0.6 (see Refs. [7578]). For higher
compositions, the Si levels become progressively deeper as the Al

0.0

0.2 0.4 0.6 0.8


Al mole fraction

1.0

Fig. 11. The composition dependence of the depth of dominant centers pinning the
Fermi level in AlGaN lms doped with Si at the density of some 1018 cm3 as
reported by four different groups: Polyakov et al. [75] (red line, open squares),
Tanyasu et al. [76] (black line, solid squares), Nakarmi et al. [77] (green line, open
circles), and Zeisel et al. [78] (blue line, solid circles). (For interpretation of the
references to color in this gure legend, the reader is referred to the web version of
this article.)

composition increases, although the actual depth of the levels


varies for various reports (see a compilation from four papers in
Fig. 11).
Correspondingly, the concentration of electrons at room
temperature also varies a lot. Part of the problem is due to the
contribution of compensating defects and impurities: carbon, VAl
O, VAlSi acceptor complexes, and oxygen DX acceptors, but
another part could be, in fact, due to Si being not a simple shallow
donor, but a deep DX acceptor. The authors of [78] have in fact
observed a high activation energy of dark conductivity of 320 meV
in Si doped AlN. They also observed strong persistent photoconductivity with an optical threshold close to 1.5 eV and the
activation energy for recovery of dark conductivity of
200  80 meV (this latter value was determined from the analysis
of PPC relaxation curves). Electron paramagnetic resonance (EPR)
spectra measurements suggested the emergence of the signal from
lled d8 states after illumination. It was proposed that the data can be
explained by the CCD of the type shown in Fig. 12 taken from Ref.
[78]. The processes involved, as described by this diagram, are the
optical ionization of one electron from the DX into a metastable state
of DX8, optical ionization of the electron on this metastable center
into the d+ donor state, capture of the electron by the d+ state with the
formation of metastable d8 state, and then thermal excitation to
return into the DX state with the capture of the second electron (the
EPR signal attributed to electrons on lled donors in Ref. [78] is
believed to be due to electrons at the d8 forming an impurity band in
which hopping is the prevalent motion mode). The ionization energy
of the d8 center is assumed to be close to 65 meV, the barrier for
capture of electron into the DX state is 200 meV. Mind, however,
that the Si ionization energy for AlN in different measurement sets
varies from very close to the hydrogen-like value of 85 meV in the
work of Tanyasu et al. [76] to 180 meV in the work of Nakarmi et al.
[77] (or Ive et al. [79]), and is the highest in the work of Zeisel et al.
[78], with the thermal ionization energy of the Si DX centers of
320 meV (Fig. 11). The extrapolated AlN ionization energy value of
Polyakov et al. [75] is also close to 300 meV.
This wide spread of the measured ionization energies can be
logically explained by recent detailed EPR spectra measurements
performed on AlN(Si) by Son et al. [80]. The authors were able to
account for the observed behavior by assuming that there is always
a competition between the DX centers due to Si and O in AlN. For
the Si DX centers the level in the gap was found to be located
78 meV below the d state. If the latter is described by the
hydrogenic model its ionization energy should be close to 60
65 meV [78], which gives for the DX state of Si in AlN the level
near Ec0.14 eV. The oxygen DX centers are deeper. Depending

10

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 12. Conguration-coordinate diagram for Si DX centers in AlN.


(After Ref. [78], Fig. 5) Copyright the American Physical Society, 2000.

on the relative concentrations of Si and O and the presence of other


compensating defects the Fermi level can be pinned at Si DX, at
oxygen DX, in between or on centers deeper than Si or oxygen
DX. The measured activation energy in dark conductivity would
then be strongly sample sensitive. For Fermi levels located high in
the gap for the Si DX states the measured activation energy will be
the average of the Si d8 and Si DX energies, i.e. close to 0.1 eV, as
measured in Ref. [76]. For deeper Fermi level the measured energy
could correspond to the Si DX state, as is perhaps the case in Refs.
[77,79]. For even deeper Fermi level the oxygen DX center will
come into play. We believe that that was what was happening for
AlN(Si) samples in Ref. [78]. The measured activation energy of the
Si DX state in that case in fact refers to the oxygen DX states.
Zeisel et al. [78] argue that this was not the case because they did
not observe the marked PPC and the d8-related EPR signal in
undoped AlN samples without intentional strong Si doping. This
argument could, however, be not so strong because their undoped
samples were very highly resistive and the Fermi level in them
could be well below the possible oxygen DX centers. Polyakov
et al. [75] showed that PPC is not detected in such cases for
undoped or Si doped n-AlGaN lms.
It should be also mentioned that in the work of Trihn et al. [81]
devoted to EPR measurements on n-AlGaN (x = 0.77) lms doped
with Si it was shown that for this composition the gap between the
DX states and d8 states of Si is very small, close to 3 meV.
Therefore, it is reasonable to assume that for lower compositions
the donors should be hydrogen-like d8 states of Si. If one looks at
Fig. 3 one can see that, for Tanyasu et al. [76] and Nakarmi et al. [77]
results the dominant donors pinning the Fermi level follow this
assumption, whereas the activation energies in Polyakov et al. [75]
and Zeisel et al. [78] are growing with composition much steeper
implying that, in these older works, the compensation by oxygen
DX acceptors and possibly by other deeper traps could be an issue
(both Tanyasu et al. [76] and Nakarmi et al. [77] claim to have
achieved reasonably low densities of oxygen and carbon on the
level of 1017 cm3).
Thus, it would seem that experimental data collected up to the
present moment suggest that Si in AlGaN forms a normal
hydrogen-like substitutional donor for Al mole fractions below
x = 0.70.8 and turns into a deep DX acceptor for higher
compositions. The activation energy of Si DX centers in AlN

should be located near 0.15 eV. For oxygen the transition occurs at
much lower Al composition x  0.3. For more Al-rich ternaries the
donors are deep DX centers, the activation energy of oxygen DX
centers in AlN should be close to 0.3 eV, the barrier for capture of
electrons close to 0.20.3 eV. The optical ionization energy is close
to 1.31.5 eV depending on composition. When analyzing the
electrical properties of AlGaN/AlN/GaN HEMTs well see that
centers with similar characteristics are often reported as present in
the HEMT barrier (see below).
Oxygen is also believed to be responsible for the high density of
residual donors in InAlN used in III-N HEMTs as an alternative
barrier material instead of AlGaN. The advantage of InAlN is that, in
contrast to AlGaN, it can be made almost lattice matched to GaN if
the In composition is xIn = 0.150.16. At the same time, high
spontaneous polarization eld at the InAlN/GaN interface creates a
very high density of two-dimensional electron gas (2DEG) close to
the density only attainable for AlN in the AlGaN system (close to
2  1013 cm2) (see e.g. Ref. [82]). However, one of the drawbacks
of InAlN is a very high density of residual donors on the order of
some 10181019 cm3 [82]. In the recent paper by Py et al. [83] it
has been shown that the residual centers are very likely due to
oxygen contamination. At that admittance spectroscopy measurements revealed the presence of two types of centers: shallow
donor-like centers D1 (activation energy 68 meV, electron capture
cross section 9.7  1017 cm2) and deep acceptor-like centers D2
(290 meV, 6.2  1015 cm2) (see Fig. 13). The former are believed
to be due to normal substitutional O donors on N site, the latter due
to DX like O centers. The authors assumed that such different
oxygen states can coexist as different structural entities e.g. due to
the compositional nonuniformities. The strong persistent photocapacitance effects observed in InAlN even at room temperature by
Py et al. [83] were attributed to the D2 DX centers. Interestingly,
when measuring the temperature dependence of conductivity in
their lms the authors observed two distinct slopes, one at low
temperatures, with the activation energy corresponding to the D1
traps, the other at high temperatures with activation energy of D2
traps. This is not a standard situation when only the DX centers
dominate: in that case the activation energy of conductivity is the
average of the activation energies of the DX-acceptor and normal
donors (see e.g. Ref. [80]).
In InGaN, both GaN-like and InN-like in the sense discussed
above in Section 2.1, the donors are always shallow, no DXlike
behavior has been observed for Si and oxygen, but, in InN-like

Fig. 13. Admittance spectra of InAlN.


(After Ref. [83], Fig. 4) Copyright American Physical Society, 2014.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

11

solutions, the main residual donors are due to hydrogen and to


native defects rather than to Si or O, as discussed in Section 2.1.
3.2. Si and O related complexes in III-Nitrides
As discussed in Section 2.1, Si and oxygen donors can form
complexes with acceptor VGa defects in n-GaN producing doubly
charged acceptors with the upper level near Ev+1.1 eV. Their
formation energy is expected to be lower than the formation
energy of triply charged VGa. In GaN DLTS spectra measurements
with optical injection (ODLTS) indeed show the presence of three
different hole traps with levels near Ev+0.92 eV [84] and Ev+(1.1
1.2 eV) [8587] that have been attributed to, respectively, (VGa
Si)2 and (VGaO)2 complexes. The former were mainly observed
in undoped and lightly Si doped MOCVD grown GaN [84], the latter
dominate in undoped n-GaN lms and crystals prepared by HVPE
and are also observed in heavily irradiated and annealed to 1000 8C
MOCVD grown n-GaN [8587]. The concentration in the as-grown
material is determined by the availability of constituent species.
Since in good quality GaN the residual donor contamination
nowadays does not exceed 1017 cm3 and the concentration of
VGa is not much higher than some 1016 cm3 the concentration of
such complexes is generally on the order of 10151016 cm3
[84]. Thus, such complexes can play a certain part in compensation
of pure undoped n-GaN and can even signicantly contribute to
rendering semi-insulating the undoped high purity GaN lms, but
it seems that deep C acceptors (see below) are usually more
efcient in this respect.
In AlGaN photoluminescence (PL) spectra measurements
revealed the presence of three groups of deep defect PL bands
that could be attributed to close donoracceptor pairs (DAP)
recombination involving shallow donors and deep acceptors [88
90]. The deep acceptor levels in each of these three groups were
approximately aligned in respect to the level of vacuum as the Al
composition and the bandgap changed [8890] (see Fig. 14 taken
from Ref. [90]). It was argued that the PL transitions series with the
lowest energy (PL peak in AlN at 3.4 eV) is due to recombination
involving shallow donors and triply negatively charged Al
vacancies VAl3. The second PL band series at higher energy
(3.9 eV in AlN) was attributed to transitions involving (VAlSi)2 or
(VAlO)2 complexes. For the highest energy series (4.7 eV in AlN)
the deep acceptor was assumed to be singly charged (VAl-2O)
complex of aluminum vacancy with two oxygen donors
[89,90]. The attribution was based on the results of theoretical
calculations and assigned the famous yellow luminescence band at
2.15 eV in n-GaN to the transitions involving the (VGaSi)2 or
(VGaO)2 complexes [8890]. This was conrmed by the doublet
splitting of respective lines caused by the O location either in the
basal plane or along the [0 0 0 1] axis of the complex [90]. Experimental results suggested that the VAl3 acceptors dominate in the
AlGaN lms with Al mole fraction higher than x = 0.58, while
vacancy-donor (not specied whether Si or O) acceptor complexes
are dominant in AlGaN lms with lower compositions, presumably
due to the difference of formation energies for various compositions. Compensation in AlGaN lms, especially for high Al
compositions, was attributed to the presence of such Al vacancy
acceptors or their complexes with donors. In AlN it has been
demonstrated that the attainable n-type conductivity of MOCVD
grown lms could be substantially increased by suppressing the
formation of VAl [89,90]. However, one has to be weary of the fact
that, as in GaN, compensation with deep C acceptors could
outweigh the effects of vacancies compensation and that the
yellow luminescence band (and related defect band series in
AlGaN) has been attributed in some recent papers to recombination on C acceptors or their complexes (we will return to this
question in Section 3.3). It should also be noted that, even for Si in

Fig. 14. The band diagram of AlGaN showing the location of different vacancy
complexes, the energies of respective PL transitions in GaN and AlN with taking into
account the doublet splitting of the two different orientations of the VIIIO complex.
(After Ref. [90], Fig. 4) Copyright American Institute of Physics, 2012.

AlGaN lms with compositions higher than x  0.7, the centers are
DX acceptors rather than simple donors (see above). This is even
more the case for O doping. Respectively, the picture of the
standard shallow donor-deep acceptor DAP recombination
becomes more complicated, especially at low temperatures (see
e.g. Ref. [91]).
3.3. Mg in III-Nitrides
Mg being the most technologically important p-type dopant in
GaN it is natural that its electronic properties have been
extensively studied. The Mg level position was established from
the temperature dependence of hole concentration and conductivity of p-GaN:Mg lms by many groups (see e.g. Refs. [9296]).
The Mg level was placed based on these measurements close to
Ev+(0.150.17) eV. There has been some published evidence on the
level becoming more shallow with increasing Mg acceptors
concentration [97]. The effect has been ascribed, as usual, to the
overlap of the individual Mg acceptor wave functions resulting in
the formation of an impurity band. Since the depth of Mg acceptors
is quite high and causing severe problems in devices sensitive to
the series resistance of the p-region, such as LEDs or LDs, changes
in the structural quality and electronic properties of GaN:Mg
occurring as the Mg concentration increases have been looked into
in some detail. It turns out that, for high Mg concentrations
approaching 102 cm3, a high density of inversion boundary
domains (IBDs), stacking faults (SFs) and dislocations is generated
due, as it is thought, to forming of Mg2N3 precipitates limiting the
Mg solubility [97101]. The high density of IBDs causes the
inversion of the dominant growth plane mode from Ga-polar at
low Mg concentration to N-polar at high Mg concentration [99],
with consequent increase in the uptake of Si and O shallow donors
and increased compensation [9799].
An additional complication comes from the hole mobility being
considerably lower than expected based on theoretical modeling.
This has been attributed to the effects of grain boundaries and the
mosaic structure of the GaN lms (see e.g. Refs. [102,103]). As a
result the low conductivity of p-GaN still remains a serious
problem in device applications and has not been completely solved
up to the present day. Using modulation doped p-AlGaN/GaN

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

12

Capacitance (pF)

superlattices (SLs) (see e.g. Refs. [104,105]) and p-GaN/GaN doped


superlattices [106,107] have been shown to provide certain
improvements in the in-plane conductivity, but the charge transfer
in the direction normal to the surface presents a problem, although
it is somewhat alleviated by the tunneling of holes between the SL
layers [105].
Mg centers in p-GaN have also been studied by admittance
spectroscopy (AS [8]) in Refs. [108114]. In general, the Mg
acceptor level positions determined from AS are qualitatively
similar to those calculated from Hall effect/conductivity measurements. However, two points have to be stressed here. First, as
discussed in Refs. [109113], the Mg case is special compared to
the case of true shallow acceptors because of the dual nature of Mg.
On the one hand, it is the most shallow acceptor determining the
samples conductivity, but on the other hand, the level is quite
deep, which could cause the electric eld build-up near the crossover point where the Mg level crosses the quasi-Fermi level [109
111]. It has also to be taken into account that, depending on the
measurement frequency, there are two distinct regimes. In the rst
the AC conductance lag is determined by the speed of holes
emission from the Mg acceptor level, in the second it is limited by
the holes delivery to the space charge region (SCR) edge, i.e. by the
Maxwell relaxation time [111]. Consequently, in the rst regime,
the measured activation energy in AS spectra and the preexponential factor are due to the Mg energy depth and the hole
capture cross section for Mg acceptors as for true deep traps in
standard AS theory [8]. In the second case, the activation energy is
that of the temperature dependence of conductivity [111]. In many
cases experimentally measured admittance spectra show two
distinct peaks in admittance/steps in capacitance [112116]. This
is illustrated by Fig. 15 for p-GaN lms grown by MBE [115], but
similar features have been observed in p-GaN lms grown by
MOCVD and HVPE [114116]. In most cases the activation energies
for both peaks are close to each other and to the Mg ionization
energy determined from Hall/van der Pauw measurements, but the
pre-exponential factor for the low temperature peak yields the
apparent hole capture cross section of 1016 to 1015 cm2, while
the high temperature peak is characterized by a very much lower
apparent hole capture cross section of 1020 to 1019 cm2 [115
118]. It has been reported that the magnitude of the capacitance
step and the admittance peak related to the low temperature
feature is greatly increased by hydrogen plasma treatment (and
proton irradiation) of p-GaN:Mg samples [115,116]. The relative
magnitude of the low temperature step was also found to be lower
in MBE growth or HVPE growth in Ar atmosphere as opposed to
MOCVD growth [115] (in both cases the hydrogen concentration is
very considerably lower than the Mg concentration, see e.g. Ref.
[119]). These observations would be consistent with the low
temperature peak being due to the MgH deep donors, as
suggested by theory [39], while the high temperature peak being
due to isolated Mg.

3000
p-GaN MBE
2000
1000
0

G/ (pF)

1000

0.18 eV,
-20
2
5.3*10 cm
0.15 eV,
-17
2
9.3*10 cm

0
50 100 150 200 250 300 350 400 450
Temperature (K)

Fig. 15. Admittance spectra of a p-GaN:Mg lm grown by MBE.

Fig. 16. Admittance spectra of p-GaN:Mg lm grown by HVPE.


(After Ref. [119], Fig. 4) Copyright American Vacuum Society, 2006.

However, in some papers it has been reported that the


magnitude of the low temperature peak increases in MOCVD
grown p-GaN:Mg upon annealing (see e.g. Ref. [112]). Also one
cannot entirely rule out that the two features in question are owing
to the two AS Mg-related regimes discussed above (i.e. limited by
the Mg interaction with the valence band and the Maxwellrelaxation-time limited effects). Clearly, more studies are necessary to clarify this issue.
It has to be noted that several reports were published in which
the Mg center ionization energy was found to be much lower than
the typical value of 0.150.18 eV (see e.g. Refs. [120122]), both in
Hall effect measurements and in AS spectra measurements. In
these cases AS spectra commonly display the normal high
temperature feature with the standard Mg ionization energy
and the anomalous low temperature feature with a low activation
energy (Fig. 16 illustrates such behavior for one such anomalous
sample grown by HVPE in Ar atmosphere [121]). In such cases the
temperature dependence of resistivity often showed clear
evidence of the prevalent hopping conductivity at low temperatures which perhaps explains the lowered measured activation
energy because the conductivity occurs by hopping of holes or
electrons along the partly lled acceptor impurity band [123]. Such
centers were mostly observed in MBE or HVPE grown p-GaN.
PL or MCL spectra of Mg doped p-GaN also present the evidence
in favor of two distinct Mg related centers. In some samples the
introduction of Mg causes the appearance of a broad blue
luminescence band near 2.9 eV, in other it gives rise to a sharp
transition near 3.2 eV with a series of phonon replicas. It was
reported that the treatment of the samples in hydrogen plasma can
convert the broad blue PL or MCL line into the sharp transition near
3.2 eV or into a mixture of both types of PL features
[115,116]. However, cases where the sharp UV line dominates
over the broad blue band in MBE samples have been reported in
contrast to the predominance of the broad blue line in MOCVD pGaN (see e.g. Fig. 17 [115]). The authors of Refs. [122,123]
suggested that the two different features observed in PL spectra of
Mg doped p-GaN lms could be attributed to the existence of two
different types of Mg centers, one of them possibly being a complex
of Mg with hydrogen. Theoretical calculations performed in Ref.
[39] explain, as already discussed above, the phenomena by the
contribution of deep donor complexes MgH with the level close to
the valence band edge. Such centers are predicted to produce a
sharp transition with a low Stokes shift as opposed to the strongly
lattice coupled transition involving isolated Mg acceptors. As
shown above the admittance spectra of p-GaN:Mg lms also bear
some evidence in favor of two types of Mg centers. But more
detailed studies are necessary in both cases to resolve the issue.

MCL 90K

0.6

10
p-MBE
1

0.1

13

Nakarmi PL
Polyakov Ea(AS) MBE
Polyakov Ea(Hall) MBE
Polyakov Ea(AS) MOCVD
Ea(Amano) Hall&PL min

0.7

p-MOCVD

Ea (Mg) (eV)

MCL intensity (Arb. units)

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

0.5
0.4
0.3
0.2
0.1

2.0

2.5
3.0
3.5
Photon energy (eV)

Fig. 17. 90 K MCL spectra in MOCVd and MBE p-GaN.

In AlGaN ternaries of various compositions, the ionization


energy of Mg has been studied by means of measurements of the
temperature dependence of hole concentration/conductivity,
admittance spectra measurements, and measurements of PL
spectra. The electrical measurements for Mg in AlGaN become
progressively more difcult with increasing the Al mole fraction
because of the increased depth of the Mg acceptor and the
progressively growing problems with compensation. Straight Hall
effect/conductivity measurements are relatively easy for Al mole
fractions x < 0.2. For higher compositions p-type conductivity has
only been reported for carefully optimized growth conditions
minimizing the impact of compensating centers among which the
prominent role of nitrogen vacancies has been noted (well return
to this point later) [124128].
In these optimized conditions, the successful p-type doping has
been reported even for AlN [97,125,128130] where prototype
LEDs have been demonstrated, albeit of rather low power and
efciency [129,130]. The Mg ionization energy determined in these
experiments was found to be close to 0.5 eV, in reasonable
agreement with theoretical predictions [39]. However, one has to
be aware of the fact that the Mg ionization energy in these
experiments has been measured for heavily Mg doped samples. At
the same time, it has been reported that the Mg ionization energy
decreases measurably for doping levels in excess of 1019 cm3
due to the formation of the Mg impurity band [97]. The
conductivity in this impurity band at low temperatures will be
of the variable range hopping Mott type that does not increase
exponentially with temperature. At higher temperatures, it will
change to the regime in which the conductivity is determined by
the carriers in the impurity band that are excited to the Fermi level,
the so called e3 type conductivity. This conductivity is increasing
exponentially with temperature, but the activation energy e3 is not
directly related to the depth of the Mg level. Rather it is determined
by the concentration of acceptors and their compensation ratio
(see e.g. Ref. [121]). Only at even higher temperatures is the
conductivity determined by the holes activation from the impurity
band into the valence band. The activation energy of this process,
e1, is the sum of the depth of the Mg acceptor level Ea(Mg) and the
e3 activation energy [32]. A comprehensive analysis of the
temperature dependences of the hole conductivity in p-AlGaN:Mg
has not been systematically done as yet. This is explained both by
the difculties with such measurements in a broad temperature
range due to the problems with ohmic contacts [125,128] and also
because of the complications caused by the holes travel over the
grain boundaries [102]. However, the existence of Mott-type and
e3-type conductivity has been shown for p-GaN grown by MOCVD
and HVPE (see e.g. Refs. [119,120]) and for p-AlGaN (x = 0.77) in
Ref. [128]).
Such ambiguities are partly alleviated when the depth of the
level is determined from admittance spectroscopy. Indeed, in that
case one deals with the holes activation into the valence band
(given, of course, that the AS peaks attribution is correctly done).

0.0

0.0

0.2 0.4 0.6 0.8


Al mole fraction

1.0

Fig. 18. Mg level depth versus Al mole fraction in p-AlGaN:Mg determined by PL


(open squares, red line), by AS in MBE (open circles, green), by Hall in MBE (open
triangles, blue), by AS in MOCVD (solid triangles, cyan), by Hall and PL (open

However, severe problems with contacts series resistance at high


Al mole fractions restrict such experiments to Al compositions
below x = 0.45. In Ref. [115] we compared the Mg activation
energies measured by the temperature dependence of conductivity
and by AS for p-AlGaN lms grown by MOCVD (composition up to
x = 0.15) and by MBE (composition up to x = 0.45). It was assumed
that the high temperature peak in admittance spectra (see Fig. 15
and the discussion above) is due to the holes exchange with the
isolated Mg acceptors (this attribution is supported by admittance
spectra measurements on p-AlGaN lms treated in hydrogen
plasma in Ref. [131] and irradiated by protons in Ref. [132]). The
results of the temperature dependence of conductivity and the AS
measurements corresponded very reasonably well, as can be seen
from Fig. 18 summarizing the results of Mg level depth
measurements performed by different groups and different
methods.
For AlGaN compositions higher than x = 0.45, it is difcult to
obtain accurate results with admittance spectroscopy. At the same
time, PL spectra measurements are readily applicable for any Al
mole fraction. Hence, these measurements were widely used to
determine the Mg level position in p-AlGaN:Mg. However, the
interpretation of the results critically depends on the attribution of
various bands observed in the spectra to certain types of optical
transitions. Fig. 19 reproduced from paper [126] shows that, in pAlGaN, a series of PL bands with energies ranging from 2.8 eV in
GaN to 4.7 eV in AlN are systematically observed. These transitions
were attributed to recombination involving Mg acceptors and
triply charged nitrogen vacancies VN3+ predicted by theory.
Growing p-AlGaN with a higher V/III components ratio (and
therefore, a reduced density of nitrogen vacancies) greatly
suppressed the intensity of such PL lines in favor of more shortwavelength lines attributed to transitions from conduction band to
Mg acceptors (respective PL bands at near 3.2 eV in GaN and at near
5.55 eV in AlN [126]). Fig. 20 taken from Ref. [126] illustrates the
proposed levels scheme in the studied p-AlGaN of various
compositions. It is suggested that growth under conditions
suppressing the formation of VN improves the p-type conductivity
in p-AlGaN and AlN due to the lower compensation ratio of Mg
[125,126]. Problem here is, of course, that, although theory does in
fact predict the predominance of VN3+ triply charged donors as the
major compensating defects in p-GaN and p-AlN [28,43], it places
respective levels in the lower half of the bandgap, near Ev+0.4 eV in
p-GaN and close to Ev+1 eV in AlN (see recent calculations in Ref.
[45] and Fig. 21 for GaN and AlN). This, of course, would be totally
inconsistent with the recombination model described in Ref. [126]
and depicted in Fig. 20. More realistically, in AlN where it is
difcult to move the Fermi level close to the valence band, one
would deal with the VN+ single donors with the charge transition
level near 1 eV from the conduction band edge (see [127] and
Fig. 22). The correspondence between this theoretically predicted
vacancy level position and experimentally determined level

14

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 21. Formation energy dependence on the Fermi level position in p-GaN:Mg for
nitrogen vacancies VN and VNMg complexes.
(After Ref. [45], Fig. 2a and b) Copyright American Institute of Physics, 2012.

Fig. 19. PL spectra in p-AlGaN as a function of Al mole fraction.


(After Ref. [126], Fig. 1) Copyright American Institute of Physics, 2009.

location is not outstandingly good, but, given possible uncertainties in determination of both values seems acceptable and thus
conrming the level scheme proposed in Ref. [126]. However, it
should be noted that the authors of Ref. [39] suggest that the
narrow near-bandedge recombination lines in p-AlGaN could be
due to the transitions involving MgH deep donors, as discussed
above. But in this case one would have to explain why growth
under more N-rich conditions in Refs. [125,126,128] greatly
improved the p-type conductivity and suppressed the longer
wavelength PL band in Al-rich p-AlGaN.
In that sense the assumption that, in AlN with the Fermi level
high above the valence band edge, the dominant compensating
states could be due not to VN3+, but to VN+ as suggested in Ref. [127]
better explains experimental observations.

Fig. 20. Levels energies and recombination scheme in p-AlGaN:Mg.


(After Ref. [126], Fig. 3) Copyright American Institute of Physics, 2009.

Another possibility could involve recombination via the levels


of VNMg complexes predicted to be formed in p-GaN [45]. They
are expected to produce a deep donor state (VNMg)2+ with the
donor level near Ev+0.85 eV [45]. In p-GaN it would be responsible
for the red PL band near 1.8 eV (see Fig. 22). If the levels in AlGaN
are aligned in respect to the level of vacuum corresponding
transition would be near Ev+1.6 eV in p-AlN and could easily be
responsible for the 4.7 eV PL band in Fig. 19. Obviously, more
studies are necessary here.
3.4. Carbon in III-Nitrides (experiment)
3.4.1. Carbon in GaN
Carbon can be intentionally introduced into GaN and other IIINitrides grown by MBE, MOCVD or HVPE by using various chemical
precursors, such as CH4 or CBr4, attempts have been made to
introduce C in GaN by ion implantation and high temperature
annealing. In MOCVD carbon can be incorporated as a result of
decomposition of group III precursors. At that, the introduction of C
is greatly facilitated by decreasing the growth pressure, as
demonstrated in Refs. [133135]. Typically the C contamination
reported for MOCVD growth was 1016 cm3 for growth at 78 Torr
and 1018 cm3 for growth at atmospheric pressure, with the C

Fig. 22. The formation energy dependence on the Fermi level position for nitrogen
vacancies (red line) in AlN grown under N-rich (a) and Al-rich (b) growth
conditions; the positive slope appearing near Ec1 eV corresponds to the
emergence of the VN+ state. (For interpretation of the references to color in this
gure legend, the reader is referred to the web version of this article.)
(After Ref. [127], Fig. 2a and b) Copyright AIP Publishing LLC, 2013.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

introduction rate monotonically increasing with pressure [133


135]. For MBE growth, the C concentration linearly grows with the
CBr4 ow and decreases with increasing the growth temperature
[136,137]. The electrical properties of C doped GaN lms indicate
the presence of relatively shallow acceptors related to carbon.
Indeed, carbon doping of GaN samples grown by MOCVD and
doped with Si led to strong conductivity compensation and high
resistivity in compensated samples [133135]. Similar effects were
observed in C doped MBE material [137139].
Both in MOCVD and MBE GaN:C samples, photocapacitance
(PC) spectra, and deep level optical spectroscopy (DLOS) invariably
point to the presence of C-related acceptor states near Ec3.28 eV
[134,137139]. Electrical measurements on C implanted and
annealed GaN lms in Ref. [140] showed the formation of an
acceptor center with ionization energy close to 0.2 eV. Similar
centers were observed in selectively regrown GaN MOCVD lms
doped with C [141]. Even shallower centers with ionization energy
close to 60 meV were ascribed to carbon for miscut semi-polar (1101) GaN lms in Ref. [141]. Acceptors shallower than Mg
observed at the interface between the semi-insulating p-AlGaN:Fe
buffer and anomalous p-AlGaN:Mg lms in Ref. [116] were
somewhat tentatively ascribed to C. In cubic GaN, carbon has been
shown to form a shallow acceptor state with a level near 0.215 eV
from the valence band edge [142,143].
These observations are in good agreement with the results of
early theoretical calculations for the position of CN acceptor level in
GaN (see e.g. Ref. [41]), but manifestly contradict the predictions of
more recent calculations in Refs. [42,43] that place the CN acceptor
level near Ev+1 eV and attribute the high resistivity in carbon
doped semi-insulating GaN to compensation of shallow donors by
these deep acceptors. If the relatively shallow acceptor centers
observed in multiple papers for GaN:C samples are not due to CN
acceptors one would expect a plausible attribution of such centers
to defects other than CN.
As already mentioned, carbon doping can be used to produce
highly resistive GaN layers by MOCVD or MBE techniques. It seems
that Tang et al. [144] were the rst to describe the effect for MBE
grown GaN lms doped with C through the addition of CBr4. The
high room temperature resistivity of the lms exceeding 108 V cm
was attributed by the authors to C compensation of shallow donors
enhancing the impact of grain boundary barriers in highdislocation density lms. The authors of Ref. [135] conrmed
the results of Tang et al. [144] and attributed the high resistivity to
self-compensation of shallow carbon acceptors, CN, and shallow
carbon donors, CGa, whose formation energies were calculated to
be close to each other for Fermi level position tending to midgap.
This, however, was expected to happen only under N-rich growth
conditions typical of MOCVD. DLOS measurements reported by
Armstrong et al. in Ref. [134] showed, however, only signals from
C-related acceptors near Ec3.28 eV, acceptors near Ec2.6 eV, and
centers near Ec1.35 eV, but no signal from CGa donors. The
Ec1.35 eV defects were attributed to carbon interstitials, Ci. The
center Ec2.6 eV was ascribed to traps with the optical threshold
near 3 eV broadened due to strong electronphonon coupling
(these centers will be henceforth called Ec3 eV) [138].
In MBE grown GaN lms doped with C or codoped with C and Si,
donor traps with levels near Ec0.11 eV associated with CGa donors
and acceptor traps Ec2.05 eV were detected in DLOS spectra in
addition to the Ec3 eV and the Ec3.28 eV acceptors. The authors
ascribed unusual hole-trap-like features in DLTS spectra of their ntype GaN (co-doped with Si and C) Schottky diodes to hole traps
near Ev+0.9 eV and associated these traps with the Ec3 eV traps in
DLOS spectra. It was demonstrated that increasing the growth
temperature in MBE not only decreases the overall C concentration
corresponding to the given CBr4 ow, but also redistributes the
concentrations of shallow and deep C centers in favor of shallow

15

acceptors Ec3.28 eV [137]. The compensation in GaN:C lms


grown by MBE, commonly under Ga-rich conditions, was
explained by the combined action of the C-related acceptors
Ev+0.9 eV and Ec3.28 eV acceptors, in contrast to MOCVD-grown
lms in which the high resistivity was supposed to be provided by
self-compensation of CN acceptors and CGa donors [135,138]. These
experiments still leave a lot to be explained. The Ec1.35 eV states
detected in DLOS spectra of MOCVD GaN:C lms have the energy
level close to the donor level of carbon interstitials Ci in recent
theoretical papers [42,43], but in these papers the formation
energy of such defects was found to be too high in n-GaN for them
to be predominant under any conditions. And it needs to be
explained why such defects are not detected in DLOS spectra of
MBE grown GaN:C lms if they are observed in MOCVD material.
The formation of CGa donors has been shown in Refs. [42,43] to be
very costly in energy and the CN-CGa self-compensation as a
possible reason for high conductivity in MOCVD GaN was
discarded on these grounds. And if the CN level in GaN is located
near Ev+1 eV, i.e. close enough to the Ev+0.9 eV observed in DLTS,
what is the shallow state Ec3.28 eV and why it becomes
predominant at high growth temperatures of GaN:C in MBE?
(Also, it would be nice to see the Ev+0.9 eV trap not as an
anomalous signal in DLTS spectra [138], but as a true hole trap in
DLTS spectra with optical injection, ODLTS; recent measurements
of ODLTS spectra in conducting MOCVD and HVPE samples and of
photoinduced current transient spectra (PICTS) in semi-insulating
samples do indeed show features that could be attributed to these
CN states [86,145], but more has to be done in terms of accurate
attribution).
All in all, it seems that a lot still has to be done to reconcile
different theoretical approaches to the problem of electronic states
of C in GaN with the results of experimental studies. This is an
important matter since carbon doping is currently widely used to
fabricate high-performance GaN-based HEMTs (see e.g. Refs. [146
148]).
3.4.2. C in AlN
Carbon is also an important impurity in AlN because it largely
determines the transparency of the AlN substrates in the UV
spectral range when using them to grow UV LED structures based
on AlGaN with high Al mole fraction. When AlN is grown by
physical vapor transport (PVT) it tends to have a high concentration of impurities, such as C, O, Si due to the crystal contamination
from the growth environment (see e.g. Refs. [149154]). Consequently, strong absorption bands appear in the crystals near 4.9 eV,
which is very detrimental to the LEDs performance. Growth of
thick AlN lms by HVPE on PVT AlN substrates is instrumental in
greatly reducing the AlN contamination (see e.g. recent papers
[155,156]) while preserving the high crystalline quality of the
material. Experiments performed in Refs. [155,156] allowed to
directly relate the absorption in the 4.9 eV band and a strong PL
band near 3.9 eV with C concentration. Theoretical calculations
placing the CN acceptor state in AlN near Ev+2 eV allowed to
attribute the absorption band to CN ionization and the PL band to
the transition from the conduction band to the CN state involving
strong lattice relaxation [156]. Additional PL band at 2.8 eV was
ascribed to the donoracceptor pairs (DAP) transition between the
CN acceptors near Ev+2 eV and singly charged VN+ deep donors
with the charge transfer level near Ec1 eV according to
calculations in Ref. [127]. The authors argue that, at high growth
temperatures and high C density, the concentrations of VN+ can be
close to the concentration of CN, particularly under the Al-rich
growth conditions. For low C concentrations the PL spectra were
dominated by the 3.5 eV PL band that is actually in a better
agreement with theoretical predictions of Ref. [90]. However, the
authors of Refs. [127,156] associated the 3.5 eV PL band with the

16

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

transitions between the shallow donors and the VAlO complexes discussed in Ref. [90], although the transition energy in
Ref. [90] was found to be 3.9 eV, i.e. close to the PL transition
claimed in Ref. [127,156] to be due to CN. The authors of Ref.
[127,156] justied their attribution by the fact that the 3.9 eV
band intensity clearly traced the C concentration, while the
oxygen concentrations in their samples were low and on the
same level. One wonders, however, if the PL bands family
described in Ref. [90] and attributed to VAlO could not be, in
fact, due to the CN transitions. For AlGaN ternary solutions it
seems that the energy positions of respective centers can be
reasonably accurately estimated by assuming that they are
pinned to the level of vacuum, so that the general dependence of
the PL peak of that kind on Al mole fraction should be quite
similar to the one reported in Ref. [90]. We here run into the
same problem as the identication of the origin of the yellow
luminescence band in GaN that has been ascribed both to
transitions involving CN and to VGaO or VGaSi transitions. It
would seem that in both cases the bands in question could be, in
fact, a combination of several bands. For GaN it seems possible
to distinguish the contributions by comparing the temperature
stability of the constituent bands [145].
3.5. Fe doping effects
Fe doping interests us mostly because of its practical importance in preparation high resistivity GaN buffer layers in GaNbased HEMTs. It seems that the rst mention of Fe doping effects in
GaN has been done in Ref. [157] where Fe was introduced into the
material grown by vapor phase epitaxy (VPE). Characteristic
intracenter PL transition near 1.3 eV was identied as belonging to
the Fe3+ state. In Ref. [158] the sharp PL lines at 1.3 eV and 1.19 eV
with ne structure due to phonon replicas observed in undoped
GaN lms were ascribed to the internal transitions in, respectively,
Fe3+ ions and Cr4+ ions. These impurities were believed to be
common contaminants coming from the growth ambience in
MOCVD of GaN. Application of Fe doping to intentionally produce
high resistivity GaN was rst reported in Ref. [159] for MOCVD
grown material and in Ref. [160] for HVPE grown GaN. In both cases
the room temperature sheet resistivity of GaN:Fe lms was above
108 V cm, the activation energy of resistivity was found to be
0.5 eV in Ref. [161]. The authors also observed the 1.3 eV PL band
due to the intracenter transitions in Fe3+ ion. Detailed studies of the
electrical properties of GaN:Fe lms prepared by MOCVD were
reported in Refs. [161164]. It was shown that the Fermi level in
these layers was pinned near Ec(0.50.6) eV level believed to be
due to the Fe induced centers. Other deep traps detected in
photoinduced current transient spectroscopy (PICTS) were electron traps with activation energies near 0.12 eV, 0.15 eV, and
0.9 eV, and hole traps with activation energy close to 0.9 eV [161
164]. The 3 mm-thick lms studied in these papers were grown in
such a way that only the 0.5 mm layer adjacent to the sapphire
substrate was heavily (up to the concentration 1019 cm3
according to SIMS [164]) doped with Fe, the rest of the lm not
doped intentionally, but still showing a relatively high density of
Fe, gradually falling down away from the substrate, but showing a
build-up to high concentration near the surface. Still, the overall
sheet resistivity was quite high, over 107 V/square, the material
showed a well dened PL 1.3 eV PL band attributable to Fe3+
internal transitions [162164]. The electrical properties were
shown to be reasonably stable upon annealing, with the decrease
of sheet resistivity by about an order of magnitude occurring only
after rapid thermal annealing (RTA) at temperatures exceeding
950 8C or after prolonged furnace annealing at 850 8C. These
changes were accompanied by the decrease in the magnitude of
the PICTS signals due to the 0.9 eV electron traps and 0.9 eV hole

traps [163,164]. The presence of the long tails in Fe doping proles


in MOCVD GaN even after the termination of intentional Fe doping
reported in Ref. [164] was also observed in Refs. [165169] and
attributed to the gas phase transport of Fe [165,168].
In GaN(Fe) lms grown by MBE and doped with Fe the extent of
these doping tails was much shorter, possibly because of a
considerably lower growth temperature in MBE experiments
[170]. However, this lower growth temperature also seems to have
led to lower Fe solubility in MBE material, so that Fe doping higher
than some 1017 cm3 was very difcult because of the formation of
twins and Fe precipitates [170].
In HVPE the Fe concentration of about 1019 cm3 could be
incorporated before serious changes in surface morphology and
increase in dislocation density started to occur [171]. In MOCVD
controlled doping with Fe at low concentrations was very difcult
because of the parasitic gas phase transport and twinning, and
crystalline quality degradation were more of a problem [168].
The question of attribution of the deep traps observed in GaN:Fe
to Fe states is not an easy one. In Refs. [172174] the excitation
spectra of the 1.3 eV line due to the internal transition in the Fe3+
ion were performed and considerably different results were
reported. The authors of Refs. [172] placed the charge transfer
(CT) energy level of the Fe3+/Fe2+ transition in GaN near Ev+3.17 eV
(Ec0.33), whereas in Refs. [173,174] the Fe3+/Fe2+ CT level was
placed at Ev+2.86 eV (Ec0.6 eV). The latter value seems to be in
better agreement with the optical absorption due to the internal
transitions in the Fe2+ ions showing a broad band peaked at
400 meV with a ne structure related to the transitions between
crystal-eld-split levels with phonon replicas [173].
Theoretical studies of Fe and Mn centers in GaN and AlN
reported in Ref. [175] seem to conrm the results of Refs.
[173,174]. Thus, the spectroscopic measurements place the CT
Fe3+/Fe2+ level in the vicinity of Ec0.6 eV, in agreement with
results reported in Ref. [160164].
Time-resolved PL spectra measurements on GaN:Fe samples
with a relatively high Fe concentration 1018 cm3 showed that
the PL relaxation time in such samples approximately linearly
decreases with increasing Fe density. These experiments allowed
to estimate the electron capture cross section of the Fe3+ state as
1.9  1015 cm2 and the hole capture cross section by the Fe2+ as
1015 cm2, thus suggesting that Fe could be an efcient recombination center in GaN [176].
At the same time, some of the published results point to
possibly deeper Fe-related states. For instance, in Ref. [177] the
Fermi level in low-pressure MOCVD grown GaN:Fe lms was
reported to be pinned near Ec1.4 eV, with prominent electron
traps at Ec1.35 eV and hole traps near Ev+0.79 eV detected by
PICTS.
The authors of Ref. [178] studied electrical properties and deep
traps spectra in bulk HVPE GaN:Fe lms. They report that, for
heavy Fe doping, the Fermi level in their crystals was pinned near
Ec0.95 eV, with the dominant deep traps in TSC and PICTS being
the traps with activation energy of 0.82 eV. For lighter doping the
dominant traps had the activation energy of 0.6 eV. In Ref. [179]
electronic states of Fe in GaN were studied by means of electron
paramagnetic resonance (EPR) and photo-EPR measurements on
HVPE GaN:Fe samples with various concentrations of shallow
donors (Si) compensated by different concentrations of Fe
acceptors. The EPR-active state is the Fe3+ state. It could be
quenched by photoexcitation with optical threshold close to 0.8 eV
that was supposed to be close to the optical ionization energy of
Fe3+/Fe2+ CT level. However, at high Fe concentrations the charge
transfer kinetics became multistage and quite complicated. Also, in
Ref. [180] it was shown that the data do not t the simple model of
shallow Si donors compensated by Fe acceptors, but requires either
that the donors and acceptors be spatially separated or that

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

additional donors be introduced with Fe, with Fe possibly creating


another state in the bandgap.
Measurements of deep traps spectra in MBE grown GaN buffers
of AlGaN/GaN HEMTs in Refs. [181,182] showed that the resistivity
of the GaN lms consisting of the heavily Fe doped layer near
sapphire and an undoped portion of the lm higher up increased
with increased Fe doping in the Fe-doped subbuffer and with
decreased thickness of the undoped portion of the buffer [181]. The
semi-insulating samples with combined buffer showed the Fermi
level pinned near Ec0.6 eV. Conducting buffers also showed the
presence of a rather high density of the 0.6 eV deep traps on the
order of 1015 cm3, but this concentration did not vary signicantly with increased Fe doping or decreased distance to the Fe-doped
subbuffer. These results strongly suggest that the compensating
Fe3+/Fe2+ states are deeper whilst the Fermi level is pinned by the
not-necessarily Fe-related Ec0.6 eV states simply because the
concentration of such states is high. The traps that are denitely
not Fe-related have been reported consistently for undoped GaN
lms, there seems to exist several types of these traps, at least one
of the traps of this family has been traced to extended defects
(partial dislocations in non-polar GaN, dislocations in GaN [183
185]).
In MOCVD grown GaN:Fe buffers of the AlGaN/GaN HEMT
structures, the two different states of that kind, one with the level
near Ec0.59 eV and electron capture cross section of
3.3  1013 cm2, the other at Ec0.6 eV and the electron capture
cross section 6.8  1015 cm2, could be clearly seen [186,187]. The
latter trap pinned the Fermi level in the semi-insulating buffer as
determined by PICTS that also showed the presence of a deeper
center near Ec0.8 eV or Ev+0.8 eV (PICTS cannot easily discriminate between the traps in the upper and lower half of the bandgap
[186]).
Studies of electrical and optical properties of GaN lms ion
doped to very high concentrations of 210 atomic percent by
implantation of transition metal (TM) impurities (Fe, Cr, Co, Mn)
suggest that the CT levels of TM could be even deeper, near
Ec(1.51.9) eV [188,189]. The TMs were implanted at elevated
temperatures of 350 8C to avoid amorphization, the samples were
annealed at 700 8C to remove as many radiation defects as possible.
Optical absorption measurements revealed the presence of very
strong absorption bands with thresholds ranging from 1.5 eV for Cr
to 1.9 eV for Mn not observed in reference samples or samples
irradiated with protons or N [188,189]. Similar absorption bands
were detected in MBE grown GaMnN and GaCrN lms with TM
concentration of over 2 atomic % [190]. The projected range of the
250 keV TM ions used for implantation was much lower than the
total lm thickness, so that conductivity versus temperature
measurements were not directly possible, but admittance spectra
measurements on Au Schottky diodes prepared on implanted
samples indicated that the Fermi level in the implanted portion
was in the upper half of the bandgap close to Ec(0.50.6) eV. Thus,
the absorption bands should have been due to transition from the
deep levels into the conduction band. It was found that the
threshold energy of these transitions were increasing in the row Cr
(Ec1.45 eV), Fe (Ec1.65 eV), Co (Ec1.7 eV), Mn (Ec1.85 eV)
[188,189]. This is the same order in which the TM levels follow
each other in GaAs, InP or GaP. Admittance spectra measurements
revealed, as already said, the important role of deep traps near
Ec(0.450.6) eV, somewhat different for different TMs, but close
to each other. The electron capture cross sections determined for
different TMs from admittance spectra at low frequencies were
also somewhat different. The conclusions arrived at in Refs.
[188,189] were that the 0.50.6 eV traps in TM doped materials
could be due to TM complexes with native defects rather than to
the TM atoms per se. It was suggested that the true Fe3+/2+ CT levels
for different impurities are located near the levels corresponding to

17

the optical thresholds mentioned above. Problem is, of course, that


the samples in Refs. [188,189] were very heavily doped and had a
high density of radiation defects not fully annealed. Even though
the optical bands in MBE grown materials were similar to the
bands observed in ion implanted samples, the concentration of TM
ions in MBE samples was very high and the introduction of TMs
promoted the formation of a very high density of shallow defects
not observed in reference material [190].
Similarly, caution is needed when analyzing the results of other
papers suggesting that the CT level Fe3+/Fe2+ lies deeper than
Ec(0.50.6) eV. For example, the states at Ec1.35 eV and
Ev+0.8 eV in Ref. [177] are suspiciously close to the levels expected
for C-related defects in low-pressure MOCVD GaN (see Section
3.4.1). For HVPE lms in Ref. [178], the observed defects are similar
to defects detected by PICTS in GaN:Fe in Refs. [161164]. One
wonders if, for changing the Fe concentration, the Fermi level could
not be shifted to one of the deeper traps not necessarily related to
Fe. Problems with interpretation of dark and photo-EPR measurements in Refs. [179,180] have already been mentioned above. Thus,
it would seem that more detailed studies are necessary to resolve
these issues and the situation could be more complicated than
proposed in Refs. [172174].
For AlGaN, an attempt to trace the changes in the Fe-related
states with increasing Al mole fraction was made in Ref. [191]. The
undoped, Si doped, Fe doped and Fe + Si doped AlGaN lms with Al
mole fractions of x = 0.15, x = 0.37, and x = 0.45 were grown by
MBE. The unintentionally doped samples of all compositions were
semi-insulating, with the Fermi level pinned by deep traps near
Ec(0.60.7) eV. Si doping with approximately 5  1017 cm3 Si
donors produced, for low Al composition (x = 0.15) lms, n-type
conductivity with the net donor density close to 5  1017 cm3. Fe
doping to the concentration of 1017 cm3 decreased the donor
density to about 4  1017 cm3 and gave rise to a deep trap in DLTS
with the level near Ec0.9 eV. It was suggested that this trap was
similar to the Ec0.6 eV trap in GaN:Fe and possibly Fe-related.
The samples with higher Al mole fraction were semi-insulating
irrespective of doping, but in the sample with x = 0.37, the Fermi
level in Fe doped and Si + Fe doped samples was pinned at
Ec0.9 eV in contrast to undoped and Si doped samples where it
was pinned near Ec0.6 eV. PICTS spectra of AlGaN(Fe, x = 0.37)
and AlGaN(Fe + Si, x = 0.37) doped samples revealed the presence
of deep traps near Ec1.1 eV, not observed in samples without Fe.
For the highest Al mole fraction x = 0.45, all types of samples were
semi-insulating, with the Fermi level pinned near Ec0.6 eV. At
that, PICTS spectra of samples with Fe doping showed prominent
peaks from traps near Ec0.6 eV (most likely, the ones pinning the
Fermi level) and near Ec1.2 eV (these latter were not observed in
samples without Fe doping). Microcathodoluminecence (MCL)
spectra consistently showed the presence of the blue MCL band
whose peak shifted to higher energy with increasing Al mole
fraction and approximately followed the change in the depth of
the Fe-related PICTS traps (0.6 eV in GaN, 0.9 eV in AlGaN
(x = 0.15), 1.1 eV (x = 0.37), and 1.2 eV (x = 0.45)) (see Fig. 23). It
was argued that these traps observed in PICTS in Fe doped samples
could be due to Fe-related deep traps similar to the Ec0.6 eV
traps in GaN [161164]. The origin of the deep states near
Ec0.6 eV in undoped samples of all compositions is not quite
clear, but comparison with presumably VN+ states observed in
AlGaN:Mg and AlGaN:O spectra in Ref. [126] indicates a similar
behavior with composition (although these states were attributed
to VN3+ in Ref. [126], Section 3.4.2). If this association of the
Ec(0.60.7 eV) with VN+ states is correct it means that, as in Refs.
[179,180], Fe doping introduces additional defects making the
situation less straightforward than the classical case of compensation of shallow donors with deep TM acceptors (see also the
discussion in Ref. [191]).

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Transition energy (eV)

18

4.5
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0

already been discussed in the previous sections, but since there


exists a certain overlap in interests well occasionally revisit
relevant issues.
In n-GaN, several prominent groups of deep centers have been
consistently observed: (1) relatively shallow traps close to the
conduction band edge, with levels near Ec(0.060.08) eV,
Ec(0.130.14) eV, Ec0.2 eV, Ec0.25 eV, Ec(0.30.4) eV; (2)
deeper traps Ec0.47 eV, Ec(0.520.62) eV, Ec(0.70.80) eV;
(3) near-midgap traps Ec(0.91.2) eV, Ec(1.281.35) eV, (4)
deep hole traps near Ev+(0.30.5) eV, Ev+(0.60.7) eV, Ev+(0.85
0.95) eV, Ev+(11.2) eV, (5) shallow hole traps near Ev+(0.1
0.2) eV. Table 1 summarizes the data on deep electron trap levels
numbered as ET1 to ET15. Hole traps HT1HT7 data are
summarized in Table 2. The notation is ours, in what follows we
correlate where possible this notation with the notation given in
original work. The spread in measured activation energies and
apparent capture cross sections of traps that one encounters in
literature is quite large which is explained by the difference in
electrical elds, in the local strain in the studied samples, the
doping level, the complicated character of the traps that are in
many cases native defects decorating extended defects. The
capture cross sections in the tables represent the apparent capture
cross sections determined from the standard DLTS procedure and
most often quoted in the literature. However, one has to recognize
that the actual cross sections could be very different from these
values (see below).
The ET1(Ec(0.060.090) eV) traps were rst reported for
electron and gamma-irradiated samples [192194] and are
believed to be nitrogen vacancy donors [192]. The ET2 and ET3
electron traps near Ec(0.120.14) eV and Ec(0.160.18) eV were
observed in gamma-irradiated [193,194], electron irradiated
[13,16,195] or proton irradiated n-GaN [196199] (see Refs.
[13,16,195] where the traps are denoted as ED1 and ED2, Refs.
[197199] where the respective traps notation is ER1 and ER2).
These centers are also generally detected in n-GaN lms grown by

Eg

Blue band
Ea(PICTS)
Ea(conductivity)
0.0

0.1

0.2

0.3

0.4

0.5

Al mole fraction
Fig. 23. The Al mole fraction dependence of the bandgap Eg for unintentionally
doped (UID) and Si doped AlGaN (solid square, red) and Fe-doped AlGaN (open
squares, red), the peak energy of the blue band for UID and Si doped samples (solid
circles, blue) and Fe doped AlGaN (open circles, blue), the activation energy Ea
(PICTS) of the major deep trap in PICTS/DLTS spectra of Fe doped samples (open
diamonds, green), the activation energy of deep traps pinning the Fermi level Ea
(conductivity) in undoped (solid diamonds, magenta) AlGaN samples. (For
interpretation of the references to color in this gure legend, the reader is
referred to the web version of this article.)

4. Other deep traps experimental studies in III-Nitrides


4.1. Deep traps in GaN
Deep traps studies in GaN have been analyzed in several
reviews touching on both the defects spectra in as-grown materials
and defects in irradiated GaN (see Refs. [1318] and references
therein). Here we would only briey recapitulate the main points
for uninitiated reader and dwell in some detail on recent results
shedding additional light on the properties and possible origin of
important deep traps in GaN and their contribution to electrical
and optical properties. Deep traps pertinent to O, C, Fe doping have

Table 1
Electron traps in undoped n-GaN, Ec  Et is the level position in respect to Ec, sn is the electron capture cross section, AS stands for admittance spectroscopy, n(T) for
temperature dependence of electron concentration.
Trap name
ET1
ET2
ET3
ET4
ET5
ET6
ET7
ET8
ET9
ET10
ET11
ET12
ET13
ET14
ET15

sn (cm2)

Et in Ec  Et (eV)

18

0.060.09
0.120.14
0.160.18
0.2
0.25
0.30.35
0.4
0.47
0.520.55
0.570.62
0.650.75
0.80.85
0.90.95
11.1
1.281.35

(2.63.1)  10
2  1016
4  1015
4  1015
(210)  1015
(310)  1016
4  1017
6  1017
(25)  1013
(26)  1015
(410)  1017
1015 to 1014
1013
1012 to 1013
1012

Method

References

Hall, AS, DLTS


DLTS
DLTS
DLTS
DLTS
DLTS, AS
DLTS
DLTS
DLTS
DLTS
DLTS
DLTS
DLTS
DLTS
DLOS, DLTS

[193,194]
[197199]
[197199]
[197199]
[13]
[208]
[206,207]
[212]
[186,187]
[184,206,207]
[206,207,219]
[17,184,220]
[17,221223]
[17,203,204,224,225]
[138,206,207,222]

Table 2
Hole traps in n-GaN and p-GaN, Et is the level position in respect to Ev, Ev + Et, sp is the hole capture cross section.
Trap name

Et in Ev + Et (eV)

sp (cm2)

Method

Possible origin

Reference

HT1
HT2
HT3
HT4
HT5
HT6
HT7

0.95
0.50.6
0.650.7
0.850.9
1.11.2
0.951.05
0.40.43

(1.72.5)  1013
2.7  1012
(1.12.3)  1014
(1.32.3)  1013
1013 to 1012
(720)  1015
1014 cm2

ODLTS
ODLTS
ODLTS
ODLTS
ODLTS
ODLTS, DLOS, DLTS
DLOS, AS

(VGa-Si)2
?
?
?
(VGaO)2
CN
CN?

[84]
[86,234,235]
[86,234,235]
[86,234,235]
[86,234,235]
[86,234,235]
[105,237]

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

MBE on sapphire [13] or on Si [200]. In the latter case a denite


correlation of the ET2 traps density with the X-ray rocking curve
half-width has been reported [200] suggesting that the traps
could be related to dislocations. The logarithmic dependence of
the DLTS peak height on the length of the injection pulse
supports the assignment of the ET2, ET3 traps with localized
centers on dislocations [197199] (this is based on the well
known work of Wosinsky in GaAs [11] that explains such
logarithmic dependence by the correlated capture by the centers
located on the dislocation line). The ET2, ET3 are the most
common radiation defects in n-GaN samples irradiated with
light particles-gamma quants, electrons, protons (see e.g. Refs.
[193199]). The ET2, ET3 traps are also commonly observed in
GaN samples after dry etching in plasma that produces nearsurface defects via low energy ions bombardment [201]. The
traps character (donor or acceptor) is still the matter of debate.
In Refs. [13,16,195] they are believed to be donors that are VNrelated. However, in Ref. [197] the traps are determined to be
acceptors based on their activation energy in DLTS being
independent of electric eld [202]. It has been noted that the
electron concentration removal rate at the early stages of
electron irradiation of n-GaN can be very reasonably explained
by the introduction of the ET3 traps which makes a serious
argument in favor of these traps being acceptors rather than
donors [203].
The ET4 Ec0.2 eV traps are quite common deep centers in
undoped n-GaN lms grown by MBE on sapphire [13] or Si [200]
(these are the N3 defects of Ref. [200]). They are also known to be
major radiation defects in n-GaN samples irradiated with electrons
or protons [13,16,197199] (in the latter case the traps notation is
ER3 [197199]). These traps are very prominent in the surface
leakage of nanopillar GaN samples prepared by dry etching
[204]. The traps density has been shown to decrease with
decreased half-width of the X-ray rocking curves in lms grown
on Si [200]. Their DLTS peak amplitude has been reported to
logarithmically depend on the injection pulse length [13,197
199,205]. Both observations are in favor of attribution of the ET4
traps to point defects decorating dislocations. Some recent
experiments to be discussed in more detail below indicate that
the ET4 defects can incorporate shallow donors as their
constituents.
The ET5 Ec0.25 eV centers are among the most common
defects in undoped n-GaN lms grown by all techniques [13,15
18]. DLTS spectra measurements on n-GaN grown by MBE whilst
varying the N to Ga ux strongly suggest that the ET5 defects
formation is favored by the N-rich growth conditions [206]. The
logarithmic dependence of the DLTS peak on the injection pulse
length provides serious grounds for assigning the ET5 centers to
nitrogen vacancies (or their complexes) decorating dislocations
[206,207].
Defects with deep levels ET6 at Ec(0.30.35) eV are
occasionally observed in as-grown undoped n-GaN lms [208],
in our own experience, often in the parts of the lms with higher
dislocation density, such as the regions adjacent to the sapphire
substrate. These centers are also seen in irradiated n-GaN
samples, but only at high doses producing very high density of
radiation defects. At that, the level of the centers grows deeper
with increasing the irradiation uence, as demonstrated by the
case of proton irradiation of n-GaN in Ref. [209]. Traps similar to
ET6 are rather commonly reported for non-polar GaN lms
grown on SiC or sapphire and having a high density of stacking
faults and dislocations. In many cases these traps are the major
centers in such lms and pin the Fermi level in them
[183,210]. The ET6 centers are also reported to be important
deep traps pinning the Fermi level at the surface of GaN lms
after dry etching. In that case they have been known to cause

19

surface leakage channels in semi-insulating GaN buffers after


etching (see e.g. Ref. [186]).
ET7 centers with levels near Ec0.4 eV are oftentimes detected
in undoped n-GaN layers grown by MBE, particularly, under N-rich
conditions [206,207]. These traps were also observed in some nonpolar multi-quantum-well (MQW) GaN/InGaN samples where
they were adjacent to the GaN/InGaN MQW region [211]. Again, as
for many other deep traps in n-GaN, DLTS peaks magnitude for
these traps has been found to logarithmically increase with the
injection pulse width [208]. These are serious grounds to attribute
the traps to native defects complexes promoted by the N-rich
growth conditions and decorating dislocations. Below well discuss
some recent experiments suggesting that the complexes in
question can incorporate shallow donors.
ET8 traps at Ec(0.470.49) eV are sometimes reported for
MOCVD grown undoped n-GaN [212]. The traps were observed to
increase in concentration when moving inside the samples grown
by epitaxial lateral overgrowth [213]. They are the rare traps for
which the normal, not logarithmic, dependence of the traps peak
magnitude in DLTS was observed [212]. The capture of electrons to
the traps required overcoming a relatively high barrier of about
0.15 eV so that the room temperature persistent photoconductivity in undoped n-GaN observed in many cases has been partly
attributed to these traps [214]. These centers were associated in
Ref. [212] with nitrogen antisite defects.
The ET9 centers with levels near Ec0.520.55 eV are encountered in MOCVD and MBE-grown undoped n-GaN lms, often
together with the deeper ET10 traps at Ec0.6 eV [186,187]. Not
much is known about the nature of these traps. They are believed
by some researchers to be complexes of Mg acceptors with
nitrogen vacancies (see e.g. Refs. [215,216]) while in other papers
they are attributed to native defects on dislocations [217] or Ferelated acceptors [218] (and also see above the section on Fe
doping), but these interpretations have not been really supported
by serious experimental evidence.
Ec0.6 eV ET10 centers are among the omnipresent defects in
undoped n-GaN grown by all techniques (see e.g. Refs. [13,17]). The
centers are very prominent in undoped non-polar n-GaN lms with
a high density of stacking faults and dislocations [26]. The
concentration of these centers in n-GaN lms prepared by NH3MBE and plasma assisted MBE was found to be enhanced by the Nrich growth conditions [206,207]. The density of the traps in ELOG
n-GaN decreases in tune with the decrease of dislocation density
with thickness [184]. The formation of the traps is suppressed by Si
doping in ELOG n-GaN [184]. The DLTS peak dependence on
injection pulse length experiments were somewhat contradictory
and will be discussed in more detail below.
Deep traps with levels near Ec0.650.75 eV, ET11, are
commonly reported for NH3-MBE and plasma-assisted MBE nGaN where they seem to be promoted by the N-rich growth
conditions [206,207]. The concentration of these traps in MBEgrown GaN increases after proton irradiation [219].
The ET12 Ec(0.80.85) eV defects are more commonly
observed in MOCVD-grown undoped n-GaN lms, their density
seems to follow the dislocation density changes with thickness in
ELOG n-GaN [184]. Neutron irradiation greatly increases the
concentration of these centers [220]. From measurements of the
electric eld dependence of the trap energy they are believed to be
donors [17,220].
The ET13 traps at Ec(0.90.95) eV are another example of very
common deep traps in n-GaN (see e.g. Refs. [13,16,17]). The traps
show a clear logarithmic dependence of the DLTS peak amplitude
on the injection pulse length and were therefore attributed to point
defect complexes decorating dislocations [13,221,222]. The traps
concentration was found to increase upon irradiation with heavy
particles (neutrons, He ions) (see e.g. [17,220,223]). For heavily

20

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

neutron irradiated n- and p-GaN the Fermi level was pinned close
to the position of these traps. From the electric eld dependence
measurements of the ionization energy the traps are considered to
be acceptors [17,224]).
The ET14 centers with apparent ionization energy 11.1 eV are
very similar to the ET13 traps in all aspects. These traps were more
commonly observed in ELOG GaN [203,224] and maskless
epitaxial overgrowth (MELO) n-GaN [225]. They are often
dominating deep traps spectra of electron irradiated samples
and nanopillar samples produced by dry etching (see e.g. Refs.
[203,204,224]).
Finally, the ET15 traps were mainly detected by DLOS (see e.g.
Refs. [138,206,207]). The traps energy quoted in Table 1 is the
optical ionization threshold energy. The traps were detected in
MOCVD-grown and MBE-grown GaN lms, and their density is
clearly related to the density of C [138]. The traps concentration
markedly increases in samples with higher dislocation density
[138] and is several times higher in MOCVD n-GaN and GaN
prepared by NH3-MBE compared to lms grown by plasmaassisted MBE [206,207] (these comparisons, naturally, were made
for similar total C concentration in all these lms [138,206,207]).
Deep traps DLTS measurements on AlGaN/GaN HEMT structures
performed in Ref. [222] report on the GaN buffer traps A3 with
ionization energy 1.3 eV observed in heavily C doped samples.
These traps could be the same traps as the ET15 traps detected in
DLOS. In that case the optical and thermal ionization energies of
such defects are quite close to each other pointing to weak lattice
relaxation effects. If this attribution is correct the ET15 traps
should also belong to defects decorating dislocations because their
DLTS peak amplitude logarithmically changes with the injection
pulse length [222], in good agreement with the trap concentration
behavior in DLOS (see above). The ET15 traps were assigned in Refs.
[138,206,207] to interstitial C donors and their energy level is
reasonably close to the level predicted by theoretical calculations
in Ref. [43].
To conclude this brief summary of the electron traps behavior in
n-GaN we would like to draw readers attention to several points.
First, as clear from the above, many of the electron traps in n-GaN
show a logarithmic dependence of the DLTS peak amplitude on the
length of the injection pulse. On these grounds the traps in
question are believed to be native defects or their complexes
decorating dislocations in GaN. However, the logarithmic dependence of the said type can stem from other factors, most notably,
from the capture within the Debye length tails in the space charge
region [226]. Careful analysis is necessary to separate this effect
from the correlated capture by centers on the dislocation line
described by Wosinski [11]. As far as we are aware, such careful
analysis has been undertaken only for the omnipresent Ec0.6 eV
ET10 traps in Ref. [227]. The authors conclusion is that, for lowdislocation-density bulk n-GaN, the traps in question are simple
point defects while in epitaxial lms on sapphire these native
defects decorate dislocations [227]. It would seem that the
logarithmic capture argument is only valid when it is supported
by the traps concentration tracing the dislocation density. Also, in
the case of centers showing the logarithmic dependence of the
trapping it is not obvious what will the apparent electron cross
section calculated from the standard DLTS analysis [1,2] correspond to and how can these data be used to determine the height of
the capture barriers from the temperature dependence of the
capture cross section. This, together with factors related to
difference in electric eld and in strain, could account for
considerable spread of activation energies and capture cross
sections of traps reported by various groups. It seems to us that the
ET6 and ET7 traps, ET8 and ET9, ET11 and ET12, ET13 and ET14
traps in Table 1 can well turn out to belong to the same or similar
defects.

To conclude this brief discussion of electron traps encountered


in n-GaN we would like to say a few words regarding the possible
composition of the above described centers and their possible role
in various phenomena in GaN-based devices. Our experiments
with electron irradiation of n-GaN lms with different donor
doping levels showed that the introduction rate of the ET2, ET3
defects considerably increased with increasing the starting donor
doping level suggesting that shallow donors could be constituent
parts of these defect complexes [203]. For the ET4 and ET7 0.2 eV
and 0.4 eV traps it was noticed that in neutron transmutation
doped GaN the latter defects determined the Fermi level pinning
position after annealing to 800 8C and the former were responsible
for the samples conductivity after annealing to 1000 8C [231]. The
concentrations of the centers obtained from admittance spectroscopy (AS) and CV proling coincided very closely with the
calculated concentration of Ge atoms produced by the nuclear
reaction of neutrons with Ga [87]. The energy level of Ge donors in
GaN is quite shallow, 19 meV [70], so that Ge donors per se could
not be the centers governing conductivity after neutron transmutation doping and annealing. Hence the close correspondence
between the Ge concentration and the concentration of the ET4
and ET7 centers was explained by us by assuming that the 0.2 eV
and 0.4 eV ET4 and ET7 traps are complexes of shallow donors and
native defects produced by neutron irradiation and annealing, such
as vacancy complexes containing various numbers of vacancies
[87].
The Fermi level pinning in heavily irradiated n- or p-GaN near
Ec1 eV was explained by us assuming that the pinning position is
between the ET12 Ec0.8 eV native donors associated with Ga
interstitials and the ET13ET14 native acceptors due to nitrogen
interstitials [228]. As already mentioned, the carrier removal rate
in electron irradiated n-GaN was, at early stages, well described by
the conductivity compensation by the introduction of ET2, ET3, ET4
traps if these traps are believed to be acceptors [203]. For higher
electron uences the carrier removal rate could be well accounted
for by the introduction of deep acceptors ET14 ascribed to
interstitial nitrogen [203]. Interestingly, a decrease in effective
dislocation density in ELOG samples compared to standard MOCVD
decreased the introduction rate of the said traps which, at rst
glance, looks puzzling [203]. The explanation of this higher
radiation tolerance of more perfect material seems to be related to
the decrease of the defect formation energy in the material with
higher starting density of defects predicted by calculations
reported in Ref. [229]. It should be mentioned, however, that, in
proton irradiated n-GaN grown by NH3-MBE, the observed carrier
removal was attributed to compensation by ET15 C interstitials
and to C-related shallow acceptors in the lower half of the bandgap
(those will be discussed in a moment, see also the C doping effects
section above) [219]. However, it needs to be understood in more
detail how this can be the case. Indeed, the centers are believed to
be due to carbon interstitials and carbon substitutional atoms on
nitrogen sites. It is not immediately obvious how can proton
irradiation increase the density of such defects. The authors refer to
the possibility that these defects are partly hydrogen passivated
and that proton irradiation removes passivating hydrogen, but it
seems more likely that protons should additionally contribute to
this hydrogen passivation if it exists? Clearly, more studies are
necessary here.
Regarding the role of deep electron traps in performance of
GaN-based devices we have already mentioned that the Ec0.6 eV
ET10 traps are believed to be the main lifetime killers in non-polar
GaN/InGaN MQW LEDs [185]. These traps also seem to play
important role in degradation of GaN-based HEMTs (see below).
The ET13, ET14 deep traps near Ec1 eV were observed to strongly
contribute to PL intensity degradation in irradiated GaN and GaN
nanopillar structures after dry etching [203,204].

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

DV bi q  N deep  w20 =2ee0;

0.0
-5.0x10

14

-1.0x10

15

-1.5x10

15

-2.0x10

15

-2.5x10

15

-3.0x10

15

-3.5x10

15

-3

2(Nd-Na)xC/C (cm )

Hole traps in n-GaN were studied by DLTS with optical injection


(ODLTS), by photoinduced current transient spectroscopy (PICTS),
by DLOS. In some cases hole-trap-like features were observed in
ordinary DLTS measurements at high temperatures and were
attributed to the operation of hole traps. Additionally, deep hole
traps below the Fermi level can produce at low temperature
persistent capacitance changes after illumination [230]. This
happens because the traps captured by deep acceptors in the
space charge region (SCR) can only be released by thermal
emission to the valence band because electrons are lacking in the
SCR. One needs to apply forward bias to supply electrons in the SCR
and restore the initial trap lling [56]. Such deep acceptors, when
lled with holes by illumination, change the apparent voltage
intercept DVbi in CV plots. The density of the traps Ndeep is related
to the change in the intercept value Vbi by the expression [230]:

21

50

sample 1
sample 2

HT2

HT3 HT4 HT1

HT5
100

150

200 250 300


Temperature (K)

350

400

450

Fig. 24. ODLTS spectra measured on two typical undoped n-GaN bulk HVPE samples
with the injection pulse of 365 nm LED, with reverse bias 2 V and time windows
100 ms/1000 ms (solid curves) and 2.5 s/25 s (dashed curves); (NdNa) in the gure
is the net donor density determined from CV measurements, C the stationary
capacitance, DC the ODLTS signal.

(2)

where w0 is the space charge region (SCR) width under


illumination (i.e., it is assumed that this is the thickness down to
which the deep traps are recharged), q is the electron charge, e0 is
the permittivity of vacuum and e is the relative permittivity. The
intercept decrease after illumination at low temperature can be
wiped out by the application of forward bias, which is the hallmark
of the effect. The traps density can be determined from Eq. (2),
whilst the traps activation energy can be estimated from the
thermally stimulated photocapacitance (TSPC) measurements
[230]. These measurements can be corroborated by ODLTS and
PICTS measurements [230232]. In the series of undoped n-GaN
lms grown by HVPE and having the dislocation density ranging
from over 109 cm2 to below 106 cm2 we observed such traps
with levels deep below the Fermi level. The density of traps varied
with dislocation density approximately as expected for the
number of broken bonds on dislocations [230,231]. Thermally
stimulated photocapacitance spectra showed that the traps in
question had the density of states peaked near Ev+0.3 eV. Similar
traps were observed in PICTS spectra of high-resistivity GaN lms
grown by MOCVD [232].
ODLTS spectra of good quality undoped and Si doped n-GaN
lms grown by MOCVD were dominated by deep hole traps HT1
with energy level near Ev+0.95 eV (see Ref. [84] where the traps are
named H1). The traps concentration increased with Si doping and
these centers were attributed in Ref. [84] to doubly negatively
charged gallium vacancy acceptor complexes with Si (VGa
Si)2. The traps concentration even in undoped n-GaN was quite
high, on the order of 1015 cm3. Therefore, these traps required
high photon ux in order to be fully recharged in ODLTS. It was
noticed that, for insufcient injection levels in ODLTS, the apparent
activation energy of the HT1 traps could be considerably higher
than the actual ionization energy that can be consistently
measured only at injection levels providing the traps saturation
[84]. The density of the HT1 traps was noticed to increase at initial
stages of electron and neutron irradiation [203,220,233]. The
thermal activation energy of the HT1 trap determined from ODLTS
measurements coincided very reasonably with the optical ionization energy of the deep acceptor responsible for the yellow
luminescence band near 2.15 eV, as determined from detailed PL
spectra versus temperature and versus excitation level measurements [14]. This suggests low impact of lattice relaxation for these
deep traps.
For thick bulk undoped n-GaN crystals grown by HVPE the hole
traps spectra observed in ODLTS were very different [86,234]. The
typical spectra are shown in Fig. 24. The HT1 hole trap dominant in
MOCVD lms could be detected only occasionally. Instead the
spectra observed with above-bandgap excitation showed the
presence of HT2, HT3, HT4, and HT5 hole traps with activation

energies 0.50.6 eV, 0.650.7 eV, 0.850.9 eV, and 1.11.2 eV,
respectively (the traps named H2, H3, H4, H5 in Refs.
[86,234,235]). The HT5 traps were absolutely dominant in various
studied samples. If one assumed that these traps are the main
compensating acceptors in undoped bulk HVPE samples there was
observed a very good linear dependence between the NdNa values
measured by CV proling and NdNa + N(HT5) values calculated
from CV proling and ODLTS (Nd here is the concentration of
residual shallow donors, Na is the concentration of all acceptors,
N(HT5) is the concentration of the HT5 centers). This linearity can
only take place if the HT5 traps are complexes involving the main
residual donors in undoped HVPE n-GaN [86].
These main residual donors are oxygen according to detailed
studies reported in Ref. [13]. Theoretical calculations discussed
above predict the existence of stable doubly charged VGa acceptor
complexes with oxygen donors (VGaO)2 with the level near
Ev+1 eV in n-GaN [22]. Thus, it makes sense to associate the HT5
hole traps with such complexes. The possibility of them being
complexes with other residual donors, Si, is ruled out by the results
of our deep traps spectra measurements on neutron transmutation
doped and annealed to 1000 8C n-GaN lms grown by MOCVD.
Before irradiation and annealing the hole traps spectra of these
MOCVD lms were dominated by the HT1 traps believed, as said
above, to be VGa complexes with Si. After irradiation and annealing
the ODLTS signal from the HT1 traps became negligible while the
HT5 traps became the dominant hole traps [87]. This correlated
very nicely with the earlier positron annihilation experiments that
demonstrated a much higher thermal stability of the (VGaO)
complexes compared to (VGaSi) complexes [236].
When measuring the ODLTS spectra of undoped HVPE n-GaN
crystals it was noticed that the signal in the high temperature
region of the spectrum where the HT5 centers were detected
decreased very rapidly when the photon energy decreased below
3 eV. At the same time the peak position measurably shifted to
slightly higher temperature indicating the presence of another
deep hole trap HT6 with a little bit lower energy Ev+(0.951.05) eV
and a much lower hole capture cross section (see Fig. 25) [86]. The
excitation spectrum of the HT5/HT6 ODLTS signal is presented in
Fig. 26 where it can be seen that there are two distinct optical
thresholds of the traps, one near 2.152.3 eV, the other near 3.1 eV.
These results were explained by us by assuming that the
concentration of the HT6 traps is considerably lower than the
HT5 traps, but, at low photon energies, the optical cross section of
HT6 center is higher and it prevails in the ODLTS spectra via direct
optical excitation of these defects. For intrinsic excitation
producing the high concentrations of free holes the HT5 centers
dominate the ODLTS spectra both because of the higher
concentration and because of the much higher hole capture cross
section [86].

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

ODLTS signal (Arb. units)

22

0
HT6

2.73 eV LED

-5
-10
HT5

-15

3.4 eV LED
-20

340

360
380
Temperature (K)

400

ODLTS signal (Arb. units)

Fig. 25. ODLTS spectra measured on one of the undoped n-GaN bulk HVPE samples
with injection pulse from LED with photon energy 2.73 eV and 3.4 eV.

10

HT5/HT6
HT4

0.1

0.01

2.0

2.5

3.0

3.5

Photon energy (eV)


Fig. 26. Excitation spectra of the ODLTS signal from traps HT5/HT6 and HT4.

Thus, it would seem that the excitation spectrum in Fig. 26


mostly refers to the optical ionization spectrum of the HT6 centers
[86]. Regarding the possible origin of these defects we refer the
reader to DLOS/DLTS spectra measurements in heavily carbon
doped n-GaN samples [138]. In the DLOS spectra the authors
observed a strong feature with the apparent optical threshold near
Ec(2.32.6) eV whose concentration obviously increased with
increasing C concentration. At the same time, in DLTS spectra
measured on Schottky diodes, an unusual hole-trap-like feature
was detected. The magnitude of this wrong sign DLTS peak
increased with C doping and the feature was assigned to C-related
deep acceptors with activation energy 0.870.98 eV and the hole
capture cross section of 7  10152  1013 cm2 [138]. This
behavior is very reminiscent of our HT6 trap. We also note that
recent theoretical calculations place the CN acceptor level near
Ev+1 eV [43]. Based on these considerations we assign the HT6
traps to the CN acceptor level [86].
The HT5 and HT6 deep hole traps are omnipresent features in
HVPE grown n-GaN, but the ratio of their concentrations strongly
varies from sample to sample [85,86,234,235]. It would seem that
detailed ODLTS spectra measurements on HVPE samples heavily
doped with Si and MOCVD samples heavily doped with C should be
very helpful in supporting the identication of the HT1, HT5, and
HT6 traps proposed above.
Carbon doped n-GaN lms studied by DLOS also showed a Crelated hole trap at Ec3.26 eV (Ev+0.14 eV) (HT7) that was
attributed by the authors to shallow CN acceptors [138]. We have
already discussed this matter in the C doping section above. There
is an obvious contradiction here with the results of recent
theoretical calculations in Ref. [43]. And there appears to be a
problem in interpreting the deep C-related centers near Ev+1 eV
attributed by theory to CN deep acceptor states. Clearly more
studies are necessary here.
It should be noted that the concentration of the major HT5 hole
traps in HVPE n-GaN lms is determined by the density of gallium
vacancies and residual oxygen donors that comprise the complex
in question. We have shown in Refs. [234,235] that by varying the

growth temperature one can obtain the optimal minimum


concentration of both constituents and obtain n-GaN lms with
extremely low residual donor concentration and very low density
of deep hole traps [235].
Another prominent hole trap in HVPE n-GaN is the H4 hole trap
with the thermal ionization energy of 0.850.9 eV. For this trap, the
excitation spectrum of ODLTS is also presented in Fig. 25. It can
be seen that, for this hole trap, the optical threshold in ODLTS is
close to 2.4 eV. These traps have never been encountered in our nGaN lms grown by MOCVD. The origin of the traps is not
understood at the moment, but they denitely include some native
point defects since it has been shown that these traps can be
produced by neutron irradiation of MOCVD and ELOG lms of nGaN [233]. The same goes for the HT2 and HT3 hole traps whose
density was measurably increased in neutron irradiated HVPE GaN
samples [233]. The HT2 traps are also often observed in n-GaN
lms and crystals grown by all techniques and heavily irradiated
with 10 MeV electrons [224]. Traps similar to HT2 centers were
reported to be produced in p-GaN after treatment in hydrogen
plasma [113] or after proton irradiation [114].
Other deep traps in p-GaN include the already described above
two Mg-related acceptor centers near Ev+0.16 eV. For MBE p-GaN
samples DLOS measurements also revealed the presence of hole
traps near Ec2.97 eV (Ev+0.43 eV) present in low density of
1016 cm3 (HT7) [237]. Similar traps were observed by us in
admittance spectra of p-AlGaN/GaN modulation doped superlattices [105]. The traps were only detected for superlattices with
inferior 2-D hole mobility and were ascribed to structural defects
in GaN lms.
To briey summarize and perhaps oversimplify: the standard
undoped MBE lms often have the shallow ET1ET4 and deep
ET10, ET11, ET14, ET15 electron traps., and H1 hole traps. The
standard undoped MOCVD layers usually have ET5, ET10, ET13,
ET14 electron traps and H1 hole traps. In HVPE lms and crystals
one often encounters ET5, ET10, ET13 electron traps. The spectra of
hole traps is considerably different from MOCVD material: one
never or seldom observes the H1 traps dominant in MOCVD, but
instead one encounters the H2H6 hole traps among which the H5
traps are commonly absolutely dominant. In general, the density of
hole traps is much higher than the density of electron traps in all
kinds of materials.
Among the hole traps in Table 2 the most important in practical
terms are, naturally, the major hole traps HT1 (Ev+0.95 eV) in
MOCVD and MBE n-GaN and the HT5 hole traps in HVPE GaN. The
former were shown to be the main agents of charge trapping in
radiation detectors on compensated MOCVD and ELOG n-GaN
[238]. The charge collection efciency of these detectors was
shown to be limited by the concentration of HT1 hole traps
[238]. Similarly, the HT5 hole traps limit the charge collection
efciency of radiation detectors built on compensated HVPE
crystals [85]. Clearly, similar effects will be important in GaNbased Schottky diode and p-i-n photodetectors.
The role of different electron and hole traps in various
photoluminescence (PL) spectra defect bands observed in GaN is
also of great importance because these defect PL bands are
commonly used to assess the crystalline and optical quality of GaN
and the bands attribution in GaN is often extended to wider
bandgap AlGaN solid solutions where PL characterization is not
uncommonly serving as the main characterization tool. Perhaps
the most famous example of defect PL bands of that sort is the
notorious yellow luminescence (YL) band in GaN. The initial
attribution of the band was to recombination in close donoracceptor pairs with donors being shallow and acceptors being deep
and having the level near Ev+1 eV (see an excellent review in Ref.
[14] and references therein). The YL band was at rst associated
with the C and O contamination, then it was ascribed to

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

The electron traps behavior in AlGaN of all compositions is


studied much less thoroughly than in GaN. It seems the rst
publication on that matter belongs to Gotz et al. [240]. The authors
reported observing an electron trap near Ec0.66 eV in lightly Si
doped MOCVD-grown AlGaN with Al concentration of 12%. It was
suggested that the trap could be similar to the very common ET5
(Ec0.25 eV) trap in n-GaN. Later Legodi et al. [241] observed
electron traps near Ec0.27 eV and Ec0.57 eV in n-AlGaN lms
with 41% Al. Park et al. [242] detected an electron trap near
Ec0.41 eV in n-AlGaN lms with 15% Al. Osaka et al. [243] studied
electron traps in HVPE-grown n-AlGaN with Al concentration of 9%
and 17%. They observed electron traps A, B, C with activation
energies 0.35 eV (0.47 eV), 0.69 eV (0.9 eV), and 0.81 eV (1 eV),
respectively (the rst value refers to the 9% Al composition, the
value in brackets is for the 17% composition). The authors note that
the energy levels of their A, B, C traps seem to be aligned in respect
to the level of vacuum. Traps near Ec0.9 eV were reported in Ref.

2000
1500
1000
500
0

Capacitance (pF)

4.2. Electron and hole traps in AlGaN

[244]. They were observed in DLTS spectra of the MBE-grown


AlGaN lms with Al concentration of 30%. In our paper [191]
already referenced above in the Fe doping section we reported the
presence of Ec0.9 eV traps in MBE-grown n-AlGaN (15% Al) lms
jointly doped with Si and Fe. For higher Al concentrations (up to
45%) the lms were semi-insulating, but PICTS data obtained for
them showed the presence of deep traps with the behavior
expected for centers with the level aligned in respect to the
vacuum level and similar in nature to the DLTS trap Ec0.9 eV in
the 15% Al lm doped with Fe and Si. MCL spectra measured on this
set of MBE AlGaN samples showed the presence of a prominent
blue band near 2.93 eV whose energy was virtually constant with
composition in the 045% Al composition range [191] (see the Fe
doping section, Fig. 23). The origin of the blue band was attributed
to electron transitions from the electron traps of the Ec0.9 eV
family to the valence band. The family itself was ascribed to the Ferelated centers, although one has to be aware that this attribution
needs more justication, as discussed above in the Fe doping
section.
In Refs. [245247] deep traps spectra of undoped HVPE-grown
AlGaN lms with 40% Al were studied by means of admittance
spectroscopy (AS), PICTS, and MCL. Admittance spectra measurements showed the dominance of relatively deep centers with
ionization energy of 0.25 eV and very low electron capture cross
section of 7  1020 cm2 (see Fig. 27), somewhat similar to the
traps reported by Legodi et al. [241]. The concentration of these
dominant traps was quite high, over 1018 cm3 from CV proling
at low frequencies [247]. They also dominated the PICTS spectra of
the samples (Fig. 28). In MCL spectra two prominent defect bands
at 3.7 eV and 2.6 eV (the latter actually comprised of two bands at
2.22.3 eV and 2.52.6 eV) were observed. Neutron irradiation led
to a rapid compensation of the dominant donors and to the Fermi
level pinning at progressively deeper centers with ionization
energies of 0.28 eV or 0.35 eV, also seen in PICTS spectra (Fig. 28).
At the highest neutron uence of 1.7  1017 cm2 an additional
deep trap with level close to 1 eV was detected. The carrier removal
rate upon neutron irradiation of such high-Al-concentration HVPE
lms was very high, around 500 cm1 [247]. Proton implantation
and hydrogen plasma treatment decreased the concentration of
deep traps, possibly due to forming complexes with hydrogen
[245,246].
The 2.2 eV and the 2.6 eV defect bands could be related to the
transitions to the levels of VIII vacancies and VIIIO complexes as
proposed by Nam et al. [88] (see above). Upon neutron and proton
irradiation the relative intensity of these bands markedly increased
at the expense of the other defect band at 3.7 eV. Increase in the
concentration of these vacancy defects could be responsible for the
observed compensation of conductivity. The 1 eV peak in PICTS of
neutron irradiated AlGaN is most likely due to these hole traps.

G/ (pF))

recombination between the VGaO deep acceptors and shallow


donors. Recently, the fashion has changed again and the band is
believed to be caused by recombination between shallow donors
and deep CN acceptors with level near Ev+1 eV or CNON complexes
[43,46,138]. It seems, however, that there are several components
giving rise to YL bands family and nding the nal solution to the
problem will always be elusive. For example, in Ref. [145] we
studied electrical properties, deep traps spectra, and PL bands in
MOCVD grown compensated GaN lms with various concentrations of C. Films with higher C concentration were semi-insulating,
lower carbon concentration resulted in compensated n-type
material. Both types of materials showed YL band in PL peaked
near 2.2 eV. However, after annealing at 800 8C the YL band
completely vanished in the conducting sample with low C density
while it remained virtually the same in the two semi-insulating
samples with higher C density. The results were explained by the
assumption that the HT1 VGaSi complexes were the dominant
deep acceptors in the sample with low C density and the deep C
acceptors were predominant in the samples with high C
concentration. The HT1 complexes can be destroyed by high
temperature annealing (see also Ref. [236]) while the C acceptors
are thermally stable [138]. For HVPE n-GaN it was long noticed that
instead of the dominant YL band common for MOCVD or MBE
grown material a prominent green luminescence (GL) band
coexisting with the YL band and the red luminescence (RL) band
were often predominant [14]. As seen in Fig. 24 the main difference
between the ODLTS spectra of HVPE and MOCVD material is the
absence or a very low concentration of the HT1 VGaSi deep hole
traps in HVPE GaN and the presence of strong HT4, HT5, and HT6
deep acceptors. When looking at the excitation spectra of
respective ODLTS signals in Fig. 26 one can conclude that the green
luminescence band can be attributed to transitions involving the
HT4 centers, the yellow band is due to the HT6 CN acceptors, and
the red band is produced by transitions involving the HT5 (VGaO)
traps. Detailed investigations of the nature of the red PL band in
HVPE GaN show that this is a donoracceptor band with the deep
acceptor level close to Ev+1.1 eV [239].
Similarly, the major broad blue luminescence band peaked near
2.8 eV in some of the p-GaN samples is often attributed to
transition involving a deep Ec(0.50.6) eV electron trap, possibly
MgVN related (see above) (see e.g. Ref. [215]). Such traps have
indeed been observed by DLOS in p-GaN [237]. However, as
discussed above in the theoretical section, the band could equally
well be explained by the strong lattice relaxation of the Mg
shallow center in GaN [39]. Clearly, more experimental studies of
both the PL bands and the deep traps in GaN are necessary.

23

800
600
400
200
0

0.25 eV, 7x10

250

300
350
Temperature (K)

-20

cm

400

Fig. 27. Admittance spectra of undoped HVPE n-AlGaN lm with Al concentration of


40%; the data are shown for several measurement frequencies of 50, 100, 200, 400,
600 kHz.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

50
15
-3
40
N=5*10 cm
30
20 N=6.7*1013 cm-3
10
0
0.63 eV
5
0.26 eV
4
3
2
1
0
50 100 150 200 250 300 350 400 450
Temperature (K)

G/ (pF)

Capacitance (pF)

24

Current (A)

10

10

-2

10

-4

10

-6

10

-8

10

-10

10

-12

10

-14

#4, 100V
0.74 eV
#8,
10V
#2, 10V,
0.7 eV
2

#1,10V, 0.25 eV
5 6 7 8 9 10 11 12
1000/T (1/K)

Fig. 29. The temperature dependence of current at voltages specied near each
curve for the PVT AlN samples from the seed (samples #1, 2), middle (sample #4),
and tail (sample #8) parts of the boule.

Hole traps near Ev+1.1 eV (Ec3.1 eV in DLOS spectra in Ref.


[244]) were also detected in DLOS spectra of MBE-grown AlGaN
with 30% Al [244]. The trap was assigned to the VAlO vacancies
and was shown to be one of the two dominant deep traps in the
studied lms, the second very prominent trap being near
Ec3.93 eV (Ev+0.16 eV) and ascribed to Mg acceptors.
As seen from this brief discussion the situation with deep traps
in AlGaN is still far from being clear. Several groups have reported
several types of deep electron traps, in each case more or less
following the vacuum level alignment model. There even exists
some overlap between certain traps detected for the ternaries with
similar Al composition, but in general different sets of data do not
correlate closely. Moreover, there are important traps that do not
seem to follow the vacuum alignment model. For example, in our
paper [191] we report that, in undoped AlGaN lms grown by MBE,
the Fermi level pinning position is close to Ec(0.60.7) eV
irrespective of the Al composition changes in the 1545% range.
Furthermore, the Fermi level is pinned near the same region of
Ec0.7 eV in high-resistivity undoped AlN crystals grown by
physical vapor transport (PVT) on SiC substrates [247]. Fig. 29
illustrates this point for several undoped AlN samples sliced from
various positions along the growth axis. In sample #2 cut close to
the SiC substrate the Fermi level is pinned near Ec0.25 eV. This
sample is conducting and the admittance spectra (Fig. 30) clearly
show the presence of two major electron traps, one with the level
Ec(0.250.28) eV, the other with the level near Ec(0.630.65) eV.
The concentrations of these centers vary for various crystals and
are within the 10131015 cm3 range for the shallower trap and
1015-mid-1016 cm3 for the deeper trap. We have argued in the
section on Si and O donors in III-Nitrides that the level of the
oxygen DX-like centers most probably is located near Ec(0.25
0.3) eV. Hence, it would stand to reason to assign the rst of

Fig. 30. Admittance spectra for one of the seed PVT undoped AlN samples.

PICTS signal (pA)

Fig. 28. PICTS spectra of the undoped n-AlGaN (40% Al) HVPE lm before neutron
irradiation (curve 1), after neutron irradiation with neutron uences of 1015 cm2
(curve 2), 3  1016 cm2 (curve 3) and 1.7  1017 cm2 (curve 4).

100
90
80
70
60
50
40
30
20
10
0

0.15 eV
0.28 eV
0.69 eV 1.28 eV

100 150 200 250 300 350 400


Temperature (K)

Fig. 31. PICTS spectra measured on PVT AlN crystal cut from the center of the boule;
the Fermi level in the sample is pinned at Ec-0.74 eV, so the 1.3 eV trap is the hole
trap at Ev+1.3 eV; measurements at 100 V, with 365 nm LED excitation, time
windows 100 ms/2000 ms (black line), 200 ms/4000 ms (red line), 300 ms/6000 ms
(green line), 400 ms/8000 ms (blue line), 500/10,000 ms (magenta line). (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)

the traps in admittance spectra in Fig. 30 to oxygen. The nature of


the second center in Fig. 30 is not clear. It pins the Fermi level in the
crystals from the middle and the tail parts of the boule (Fig. 29). It is
also clearly seen in PICTS spectra of the semi-insulating crystals
with the Fermi level pinned near Ec0.7 eV (Fig. 31). But whether
these traps are similar to the traps pinning the Fermi level in
undoped AlGaN lms grown by MBE is not clear. If so these are
strange centers whose level is located approximately in the same
place of the bandgap and is not tied to the level of vacuum for a
wide range of Al compositions?
The PICTS spectra in Fig. 31 show in addition to the Ec0.7 eV
electron trap a more shallow electron trap near Ec0.15 eV (very
likely due to Si, see the discussion in respective section above) and
the hole trap at Ev+1.3 eV (since the Fermi level is pinned at
Ec0.7 eV, all traps with higher energy should be hole traps when
the measurements are performed on samples with ohmic contacts,
as in this case). The nature of these hole traps is not quite
understood. They are too shallow to be ascribed to triply negatively
charged Al vacancy acceptors VAl3 or doubly negatively charged
acceptor complexes (VAlO)2 acceptors if one assumes the PL
recombination model described in Refs. [88,90]. Following Ref. [89]
the trap we observe could be in principle attributed to the singly
charged acceptor complexes (VAl2O) with acceptor level in the
vicinity of Ev+1.3 eV and respective PL band near 4.7 eV. If that
were so, the blue luminescence band in GaN near 2.9 eV could be
due to a similar acceptor complex (VGa2O) located near Ev+(0.5
0.6) eV and could be associated with the HT2 hole traps described
above.
MCL spectra in our crystals indeed showed a prominent 4.4 eV
band that could be, with some stretch, associated with the 4.7 eV
band predicted for the (VAl2O) acceptor complexes of Ref. [89],
but only for samples cut from the middle of the boules. For the
samples cut from the seed part the absolutely dominant band was

MCL intensity (Arb. units)

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

25

5. Deep traps in AlGaN/GaN and in GaN-based HEMTs


10

5.1. AlGaN/GaN heterojunctions: the origin of two-dimensional gas


and deep traps in the material

middle
1
seed
0.1
2

4
5
6
Photon energy (eV)

Fig. 32. Room temperature MCL spectra measured on PVT AlN samples cut from the
seed and middle parts of the boule.

the 3.4 eV band attributed in Ref. [88] to transitions involving the


triply negatively charged VAl3 acceptors (this attribution is
supported by the results presented in Ref. [248]). The 3.4 eV band
was also clearly visible in the spectra of the samples cut from the
middle of the boules (Fig. 32). Interestingly, the defects responsible
for the 3.4 eV MCL band (VAl3?) were shown to decorate the lowangle dislocation boundaries in AlN [247] (Fig. 33). Thus, the
observed data on hole traps in GaN and AlN (and AlGaN) is
reasonably tied together, although not totally without questions.
However, this idyl is not complete, unfortunately. As mentioned
in the previous sections, alternative PL bands attributions
involving CN acceptors have been proposed for AlN [156] and
nd some grounds in the results of recent theoretical calculations
[43,156]. Then the alternative interpretation for the 1.3 eV hole
traps in AlN could be as due to the donor state of the CN center
predicted in Ref. [156]. But obviously a lot more experimental and
theoretical work is required to clarify these issues.

Fig. 33. MCL image of the surface of one of the PVT-grown AlN samples obtained for
the 3.4 eV MCL band registration mode and showing the decoration of grain
boundaries by defects producing these bands (VAl3).

The most prominent feature of AlGaN/GaN heterojunctions is


the fact that the two-dimensional electron gas (2DEG) of very high
density on the order of 1013 cm2 can be formed without special
donor doping of the AlGaN barrier, in contrast to the heterojunctions in the AlGaAs/GaAs system. Thus, even given the considerably lower 2DEG mobilities in the AlGaN/GaN system, the channel
conductivity in high electron mobility transistors (HEMTs) is
comparable or higher than in AlGaAs/GaAs, while higher conduction band offsets and higher electron saturation velocities in
AlGaN/GaN together with very high electrical breakdown elds
make these heterojunctions very attractive for high-power and
high-frequency applications. The origin of this polarization
doping can be traced to the high spontaneous polarization and
piezoelectric polarization elds in wurtzite III-Nitrides lacking the
inversion symmetry. For the Ga polarity growth in the [0 0 0 1]
direction (predominant in HEMT applications) both polarization
elds in AlGaN/GaN hetrojunctions point toward the AlGaN/GaN
interface [249,250]. This should result in the formation of the
positive xed ionic sheet charge at the interface and a negative
sheet charge at the surface balancing the positive charge. This in
itself does not create the 2DEG at the heterointerface. For that to
happen there should be ionizable donors in the barrier that can
contribute their electrons to compensate the positive sheet charge
at the AlGaN/GaN interface. Hsu and Walukiewicz [251] argued
that the presence of amphoteric surface states with the charge
transfer (CT) level  somewhere near Ec1 eV could simultaneously
pin the Fermi level on the surface in the vicinity of the CT level and
render the surface traps to be positively charged. Thus, such states
contribute to the compensation of the polarization positive charge at
the interface, downwards band bending at the interface on the GaN
side, and the formation of 2DEG. In this model the 2DEG density is the
function of the AlGaN/GaN conduction band offset DEc and the
thickness of the AlGaN barrier. This model was further developed in
Refs. [252,253] to formulate the so called surface donor concept in
which it is assumed that the surface donor with the level Ec  ED is
providing the electrons to the 2DEG. The 2DEG sheet density ns
increases with increasing the density of polarization charge sint (and
hence the Al mole fraction) and with the increasing thickness of the
AlGaN barrier. The latter factor is due to the band bending in AlGaN
regulating the surface donor lling with electrons. If the surface donor
level is always below the Fermi level it should be lled and cannot
contribute to the formation of 2DEG; for the given surface donor level
position ED there is the critical thickness dcr at which the surface
donor is ionized at a given Al composition. The 2DEG density
increases with increasing barrier thickness until the entire interfacial
polarization charge is compensated. The basic equations of the model
are [252]:
dcr ED  DEc eAlGaN =qs int ;

(3)

qns s int 1  dcr =dAlGaN ;

(4)

where q is the electron charge, eAlGaN is the dielectric permittivity


of the barrier. By tting the experimentally observed dependences
of ns on the barrier thickness the ED value was calculated to be
Ec1.65 eV for MOCVD grown AlGaN/GaN heterojunctions with
xAl = 0.34 [252] and Ec1.42 eV for MBE grown AlGaN/GaN with
xAl = 0.27 [253].
When the 2DEG is being formed under the Schottky barrier the
electrons to compensate the interfacial polarization charge are
provided by the interface traps at the metal/AlGaN boundary and

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

26

the Schottky barrier height Fb plays the role of the surface donor
level [254]:
ns s int =q  eAlGaN qFb  DEc EF =q2 dAlGaN 
 Nt1 dAlGaN =2  Nt2 wGaN =2;

(5)

where EF is the Fermi level position in respect to the GaN


conduction band, Nt1 is the density of deep traps with level above
the Fermi level in the AlGaN barrier, Nt2 is the density of deep traps
in the GaN buffer with the level above the Fermi level in the buffer,
wGaN is the thickness of the space charge region in GaN buffer
adjacent to the heterointerface.
The treatment above neglects the fact that there are excited
states in the quantum well at the AlGaN/GaN heterointerface that
can be partially lled. The full model should be based on
simultaneous solving of the Poisson and Schroedinger equations,
but what makes this modeling possible is the fundamental idea
that the entire polarization charge can be compensated by
ionization of the surface states in the barrier. More generally, all
states in the barrier that have the levels above the Fermi level
contribute to the 2DEG density formation. The modeling is carried
out using the published data on polarization elds in AlGaN and
GaN of various compositions, the Schottky barrier heights
dependence on Al mole fraction, the known dependences on
composition of the lattice parameters, the bandgap values and the
band offsets values, the elastic moduli and the spontaneous and
piezoelectric polarization values dependence on the Al mole
fraction (see e.g. the general description in Ref. [255]; the relevant
parameters can be found e.g. in Refs. [249,250]). Multiple modeling
packages have been developed and are to be found on the market
or in the literature (see e.g. [256]).
Understandably, these modeling packages as a rule neglect the
localized traps in the barrier and in the GaN buffer while the
nonideal behavior of AlGaN/GaN HEMTs is determined by trapping
at the surface, in the AlGaN barrier, and in the GaN buffer and needs
to be studied experimentally. In this section we discuss such
measurements for heterojunctions, the next section deals with
trapping effects in the actual HEMT devices. Currently it is widely
recognized that traps in various parts of HEMT structures are
causing parasitic effects, such as gate and drain lags, frequency
dispersion, breakdown at voltages lower than theoretically
expected, and are closely related to degradation of transistor
parameters. Some of the traps in question belong to the realm of
device fabrication and can be eliminated by optimizing device
technology, for example, introducing surface passivation routines,
depositing gate dielectric layers, varying the Schottky gate
deposition technology, dry etching procedure, etc. Others, such
as the traps in the barrier, in the buffer, at the AlGaN/GaN
heterointerface, the type and density of extended defects and
surface defects, are the characteristics of the growth process.
Ideally, one would like to determine from deep traps analysis
performed on simple test structures a measure of the suitability of
the grown epitaxial structures for fabricating well behaving HEMTs
without going through the complicated procedure of building and
characterizing the entire devices. The simplest test devices are, of
course, plain heterojunctions with ohmic contacts and Schottky
diodes. These allow in principle to determine the most basic
characteristics of the heterostructures: the 2DEG concentration
and mobility, the threshold voltage for depletion of 2DEG, the
presence of deep traps from conventional deep traps analysis
procedures, the structural perfection and its relation to electrical
properties and deep traps.
Experimental measurements on AlGaN/GaN heterojunctions
indicate that, in fact, the density of deep traps in the barrier, at the
interface, and in GaN can be quite substantial to measurably affect
the heterojunction characteristics. In early studies it has been

reported that strong persistent photocurrent can be produced in


AlGaN/GaN heterojunctions. The persistent photocurrent IPPC(t)
relaxation kinetics at room temperature could be described by the
b
stretched exponents of the form IPPC(t) = IPPC(0)exp[(t/t) ]. The
4
characteristic relaxation time t was on the order of 10 s at room
temperature. Measurements of the spectral dependence of such
photocurrent indicated a signicant role of states near Ec3.36 eV
in the AlGaN (x = 0.15) with concentration comparable to the 2DEG
density [257].
In Ref. [258] very long persistent photocurrent transients
described by stretched exponents with characteristic relaxation
time t of some 104 s at 305 K were observed in AlGaN/GaN
heterostructures (x = 0.3). The spectral threshold for the effect was
close to 1.85 eV and the concentration of PPC electrons was
comparable to 2DEG density. Treatment in CF4 plasma greatly
increased the magnitude of the traps giving rise to the PPC
photoconductivity with optical threshold near 1.85 eV (Fig. 32) and
strongly shifted the threshold voltage corresponding to 2DEG
depletion toward positive voltages. The effect was explained by the
introduction of negatively charged F ions into the AlGaN barrier,
but the similarity of optical thresholds in the untreated and Ftreated structures suggests that, in fact, F treatment could simply
increase the density of deep traps in the barrier already existing
before treatment (CF4 treatment of that kind is one of the
established ways to fabricate enhancement-mode or normally
closed AlGaN/GaN HEMTs [259,260]). The temperature dependence of the characteristic PPC lifetime t produced the activation
energy of 0.62 eV, much lower than the optical threshold in Fig. 34.
Our group has reported that in MOCVD grown AlGaN/AlN/GaN
HEMT structures (x = 0.2, 0.3, 0.4, 0.5) with Ni Schottky barriers the
threshold voltage corresponding to the depletion of 2DEG was
considerably lower (less negative) than predicted by the theoretical model neglecting possible impact of deep traps [261] (see
Fig. 35(a) and (b)). This was explained by the presence of additional
negative charge due to deep traps in the barrier, at the interface
with GaN or in GaN. The shift of the threshold voltage DVth can be
converted into the barrier/interface charge Qi if the excessive
charge distribution is known. For the charge sheet at the surface it
would be [262]:
Q i DV th C acc ;

(6)

(Cacc here is the capacitance in accumulation); for uniform


charge distribution in the barrier the charge in (6) should be
divided by 2 [262]). The results in Fig. 35 indicate the presence of

Fig. 34. Persistent photocurrent spectra at 305 K in untreated and F-treated AlGaN/
GaN, the inset shows the actual photocurrent relaxation curves tted with
stretched exponents.
(After [258], Fig. 2) Reprinted with permission, Copyright 2008 Wiley-VCH Verlag &
Co).

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Capacitance (pF)

1000

27

PPC&
PPC+3V

30% Al at 85K
V(PPC-dark)=0.7V
-9
Q=V*C=1.4*10 C
12
-2
Nss=1.7*10 cm

100
0V
10

-4.1V

-3.7V
1

-4

-2

Voltage (V)
Fig. 37. 85 K CV characteristics of the AlGaN/AlN/GaN (x = 0.3) structure measured
after cooling in the dark at 0 V (black solid line), after cooling in the dark at 3.7 V
(blue line, dashed), at 4.1 V (cyan line, dash-dot), after cooling at 4.1 V in the dark,
illumination at 85 K and 15 min wait (red line, dotted, marked PPC, i.e. persistent
photocapacitance), and after additional application of +3 V of forward voltage
(magenta dashed line marked PPC + 3V). (For interpretation of the references to
color in this gure legend, the reader is referred to the web version of this article.)

Fig. 35. (a) Experimentally measured CV characteristics of a set of AlGaN/AlN/GaN


structures with various Al mole fractions in AlGaN, (b) CV characteristics
calculated theoretically.

V(PPC) (V)

2.0
50%

1.5
1.0

40%
0.5
30%

-2

2.5x10

12

2.0x10

12

1.5x10

12

1.0x10

12

20

25

30

35

40

45

50

3.5

Capacitance (pF)

-5.4V PPC

100
10
Dark

1
100

-5.4V PPC

10
1
Dark
0.1
50 100 150 200 250 300 350 400 450
Temperature (K)

Fig. 39. The temperature dependence of dark and PPC capacitance C and AC
conductance divided by circular frequency G/v for partial depletion.

higher at low temperature of 85 K than at 400 K when the cooling


down took place at 0 V (see Fig. 40).
This tunneling additionally charges the traps in the barrier
when the sample is cooled at high reverse bias and causes the
positive voltage shift of CV characteristics (Fig. 37) and the strong

2
12

3.0

1000

Current density (A/cm )

3.0x10
Nss (cm )

12

2.5

Fig. 38. The spectral dependence of the PPC threshold voltage shift.

10
3.5x10

2.0

Photon energy (eV)

G/ (pF)

some 1012 cm2 charged acceptors in the barrier, with the density
of acceptors increasing with Al mole fraction (Fig. 36).
It was also observed that the threshold voltage in CV
characteristics measured at low temperature strongly depended
on cooling conditions: the threshold voltage was more negative for
cooling at 0 V then for cooling at the negative bias close to
threshold voltage, presumably owing to negative charging of deep
traps (Fig. 37). Illumination at low temperature shifted the
threshold voltage back toward the more negative voltages due
to the optical ionization of the traps charged during cooling down
at high reverse voltage (Fig. 37). Fig. 38 presents the spectral
dependence of the effect and shows the presence of the optical
threshold near 1.8 eV. The shift after illumination was persistent at
low temperatures. Heating the sample up to temperatures above
room temperature was required to return to the starting
conditions (see the dark and persistent capacitance and conductance temperature dependences in Fig. 39). It was suggested that
the reason for the observed voltage shifts was the tunneling of
electrons from the Schottky metal into deep states. Indeed, the
reverse IV characteristics of all studied diodes showed the
prevalent tunneling via deep traps manifested in very weak
temperature dependence of reverse current and in the saturation
of the reverse current at reverse voltage close to the threshold
voltage. In fact, the current at high reverse voltages was even

0.0
1.5

10

-2

10

-4

10

-6

10

-8

10

-10

10

-12

85K, 0V
290K
400K

85K, -4V
-4

-3

-2

-1

Voltage (V)

Al composition (%)
Fig. 36. Estimated deep trap densities calculated from the measured threshold
voltages shifts compared to theory.

Fig. 40. IV characteristics measured at different temperatures and different


voltages (marked near the lines) for the AlGaN/AlN/GaN (x = 30%) structure, mind a
strong decrease of reverse current at 85 K when cooled at high reverse bias.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

28

decrease of the reverse current at low temperature (Fig. 40). The


latter effect is due to the lling of the traps participating in
tunneling. Illumination removes the additional negative charge
shifting the threshold voltage back to more negative values.
Interestingly, application of forward bias then cannot recharge the
barrier traps indicating the presence of a barrier for capture of
electrons. These processes should occur under the Schottky diode,
not on the free surface because the electrons tunneling from
Schottky diode do not travel distances exceeding 0.51 mm from
the Schottky diode edge as clearly demonstrated by Kelvin probe
surface potential scanning e.g. in Ref. [263].
The results could in principle be interpreted as the decrease of
the positive charge in the barrier/interface/buffer, but the presence
of the barrier for recapture of electrons removed from the centers
by illumination at low temperature rather suggests the  charge
transfer of DX-type or the presence of a barrier spatially separating
emitted electrons from the capture site.
The thermal activation energy and electron capture cross
section of the traps responsible for the temperature metastabilities
described by Figs. 3740 can in many cases be determined by the
reverse DLTS method in which the partially depleted heterostructure is pulsed to high reverse voltage to ll the traps in
question and then returned to the quiescent bias
[181,182,261,264] (the procedure is similar to lling the traps in
HEMTs by reverse bias pulsing to determine parameters of the
traps in HEMTs responsible for the gate lag or drain lag, see next
section). The capacitance relaxations are then processed in the
usual DLTS manner.
Fig. 41 shows the example of such spectra measured on the
AlGaN/AlN/GaN structures with different Al mole fractions in the
AlGaN layer. It can be seen that deep traps with activation energies
0.65 eV (x = 0.3), 0.7 eV (x = 0.4) and 0.85 eV (x = 0.5) dominate
such reverse DLTS spectra of AlGaN/AlN/GaN structures.
The marked difference between the optical ionization energy
and the thermal ionization energy of the centers points to the
strong lattice relaxation upon electron capture/emission. (Also
mind that the thermal activation energy of the centers in question
is close to the activation energy of PPC relaxation times in Ref.
[258]).
Centers of the above described type are quite common for
variously grown GaN-based heterostructures and have been
observed in AlGaN/GaN heterostructures grown by MOCVD on
Si, SiC, and on sapphire [265267], in addition to the above
discussed AlGaN/AlN/GaN structures on sapphire. We detected
centers with similar behavior in MOCVD grown InAlN heterojunctions [268]. They were also found to increase in concentration after
neutron [268] or electron [269] irradiation. These data suggest the
participation of major technologically important contaminants
(most likely, O, C) forming complexes with native defects created
by irradiation. At the same time, in AlGaN/GaN, AlN/GaN with

semi-insulating GaN(Fe) buffers grown by MBE [181,182,264] the


centers dominant in reverse DLTS were the 0.25 and 0.6 eV
prevalent in the GaN buffer.
Standard DLTS measurements on the samples with large area
(diameter 1 mm) Schottky diodes were very difcult to interpret
because of the problems with high series resistance giving rise to
peaks of the wrong sign. Also, the shift of the steady-state CV
characteristics of the diodes caused by the metastabilities
described above presents serious problems. Using concentric
Schottky diode/ohmic contacts structures with a low distance
between the Schottky diode and the ohmic contact alleviates the
problem and allows to measure the standard DLTS spectra when
the metastable shift of the CV characteristics with temperature is
not strong (see e.g. Ref. [186]).
However, even when the roll-off frequency of the diode
capacitance was higher than the probing frequency in DLTS
(typically 1 MHz) one has to pay attention to the capacitance
versus frequency Cf characteristics under high reverse voltage
depleting the 2DEG, particularly for the case of semi-insulating
buffers. Fig. 42 illustrates such a case for the MOCVD grown
concentric AlGaN/GaN structure on semi-insulating GaN(Fe)
buffer. It can be seen that, for high reverse voltages that are of
interest for probing the different layers of the structure, the roll-off
frequency becomes progressively lower than 1 MHz with increasing reverse bias. The reason for that is the undercutting the 2DEG
portion of the channel below the Schottky diode and thus
increasing the effects of series resistance. In that sense multinger
rectangular Schottky diodes with narrow ngers are preferable
(see [265,267,269,270]).
Several groups have reported standard DLTS measurements on
AlGaN/GaN heterostructures. There is no mention in these papers
on strong variations with temperature and applied bias of the CV
characteristics, as discussed above. One has to assume that in these
cases the density of the barrier/interface traps of the type
illustrated by Figs. 3740 was not so strong. The authors of Ref.
[221] measured DLTS spectra in AlGaN/GaN heterojunctions
grown on SiC substrates and either having or not having SiNx
passivation of the free surface. They reported the dominant role of
two electron traps A1 (activation energy close to 1 eV) and A2
(activation energy 1.2 eV). The A1 traps had a very high electron
capture cross section of 2  1012 cm2. They showed a logarithmic
dependence of the peak amplitude on the injection pulse length
characteristic of extended defects [221]. The traps in question were
similar to the traps observed by the authors in n-GaN, particularly
electron irradiated GaN (see the section above on deep traps in
GaN). On the strength of these arguments the A1 traps were
associated with nitrogen interstitials acceptors decorating dislocations in GaN. At high temperatures corresponding to the
emergence of the A2 traps hole-trap-like sign of the signal was
often detected. The authors did not observe a strong inuence of

25

Reverse DLTS for HEMT30&40&50


0.25 eV

20
15

0
50

0.85 eV
0.7 eV
40%

10
5

0.6 eV

50%

0.65 eV
30%

100 150 200 250 300 350 400

Capacitance (pF)

RDLTS signal (Arb. units)

1000
30

0V -2V -2.5V -2.7V

800
600

-2.8V
400
200
0
10

-1

10

10

10

10

10

Frequency (kHz)

Temperature (K)
Fig. 41. Reverse DLTS (RDLTS) spectra measured in AlGaN/AlN/GaN structures with
different Al mole fractions in the AlGaN barrier.

Fig. 42. Capacitance versus frequency dependences for MOCVD-grown AlGaN/


GaN(Fe) sample measured at room temperature for different biases (marked near
the curves).

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 44. DLTS spectra measured on AlGaN/GaN MOCVD grown structure.


(After Ref. [271], Fig. 3) Reprinted with permission, 2010 Wiley-VCH Verlag
GmbH & Co.

AlGaN/GaN. The authors observed a measurable hysteresis of


CV characteristics at low temperatures, but no strong shift of the
threshold voltage of the type described by Figs. 3740. The DLTS
spectrum reported in Ref. [271] is shown in Fig. 44. The energies of
the 6 traps A, B, C, D, E, F were respectively 0.15, 0.21, 0.12, 0.42,
0.49, 0.94 eV. The D, E, F traps were believed to be located in the
buffer or at the AlGaN/GaN interface. Measurements of the
injection pulse length dependence of the peaks amplitude
revealed the co-existence of point-defect-related and extendeddefects-related behavior suggesting that the observed peaks
corresponded to isolated defects or defects decorating dislocations, both species having closely spaced defect levels.
In Ref. [272] DLTS measurements revealed the existence of
Ec0.3 eV electron traps in the buffer of AlGaN/GaN HEMT
structures. Strong hole-trap-like peaks with activation energy
0.8 eV were reported for AlGaN/GaN DLTS spectra measurements
in Ref. [273].
For MOCVD grown AlGaN/GaN structures with semi-insulating
GaN:Fe buffer, we reported the presence of two electron traps with
activation energies of 0.59 eV and 0.6 eV (Fig. 45). The 0.6 eV trap
was similar to the trap pinning the Fermi level in the GaN:Fe buffer
and could be associated with the main center in the buffer.
However, the high sheet density of these traps on the order of
1012 cm2 suggests that we rather observed the accumulation of
such traps near the AlGaN/GaN interface [187]. This accumulation
was more prominent after oxygen annealing at 750800 8C [186]
(see Fig. 45).
It should be noted that the hole-trap-like signals detected in
several DLTS studies described above (and in current or capacitance DLTS measurements on transistors, see next section) are not

DLTS signal C/C (%)

passivation on measured spectra and indeed on the overall


electrical performance of their diodes. In a more recent paper
the deep traps spectra in AlGaN/GaN heterojunctions grown on
semi-insulating GaN:C buffers on SiC were obtained by standard
DLTS measurements [222]. It was found that the behavior was
somewhat different for buffers with high C concentration (grown
at low pressure, see the section on C in GaN above) or low C
concentration (grown at atmospheric pressure). For buffers with
high C concentration three electron traps A1, A2, A3 were observed.
The A1 and A2 traps were similar to the ones detected in Ref.
[221]. The A3 trap had the activation energy of 1.3 eV. All traps
demonstrated logarithmic dependence on the injection pulse
length. The magnitude of respective peaks was increasing for
reverse biases and pulsing conditions probing the region closer to
the GaN buffer. At high temperatures the authors also observed a
hole-trap-like peak H2 with the activation energy 1.3 eV believed
to be located closer to the AlGaN/GaN interface. With low C
concentration in the buffer the A1, A2 electron traps were still
observed. In addition an Ax electron trap with the activation energy
of 0.9 eV was present near the buffer. At high temperatures over
450500 K the hole-trap-like peaks H3 with the activation energy
of 1.24 eV were commonly detected. These hole traps were more
prominent for the space charge region boundary located closer to
the interface (the ndings are illustrated by Fig. 43 from Ref. [222]).
In Ref. [271] DLTS measurements were performed on Schottky
diodes of test FET structures prepared on MOCVD-grown

Fig. 43. DLTS spectra observed in AlGaN/GaN(C) structures with the low C
concentration in the buffer (upper panel) and high C concentration in the buffer
(lower panel); the spectra were measured at a xed reverse bias and changing
height of the injection pulse (the actual values are presented in the gure labels).
(After Ref. [222], Fig. 3) Reprinted with permission, Copyright AIP, 2010.

29

50
45
40
35
30
25
20
15
10
5
0

0.6 eV,
-15
2
6.8x10 cm
0.59 eV,
-13
2
3.3x10 cm

100 150 200 250 300 350 400


Temperature (K)

Fig. 45. DLTS spectra measured on MOCVD grown AlGaN/GaN(Fe) heterostructures.

a trivial occurrence. Pulsing in Schottky diodes should only provide


electron lling and, for traps above or close to the Fermi level,
should produce electron-trap-like signals (capacitance increasing
with time after the end of the injection pulse). To observe the hole
trap peaks one should recharge the traps below the Fermi level
with holes. For majority-carrier devices, such as Schottky diodes or
HEMTs this should not happen. However, with strong electric eld
at high reverse bias and prominent tunneling one can expect the
electrons from the traps in the lower half of the AlGaN bandgap to
tunnel into the states in the GaN buffer, particularly if the buffer is
semi-insulating, as illustrated by the band-bending diagrams in
Fig. 46 taken from Ref. [237]. Then the traps in the AlGaN barrier
will return to the equilibrium by thermally emitting holes thus
producing hole-trap-like signals. However, one has to be weary of
effects of the series resistance that can easily come into play and
can contribute to the appearance of the hole-traps-like signal (see
Fig. 42). For example, in Ref. [186] we observed normal electrontraps-like DLTS spectra for our concentric-ring Schottky/ohmic
arrangement only after oxygen treatment that signicantly
decreased the series resistance due to the access region of 2DEG
[186] (see Fig. 47). For virgin structures with higher series
resistance the spectrum was strongly affected by the negative
hole-trap-like signal due to the series resistanse effects. In order to
check, whether the hole-trap-like signals in DLTS spectra are
indeed due to real hole traps ionized by tunneling in the presence
of high eld it is useful to measure DLTS spectra with optical
excitation (ODLTS) or do the deep levels optical spectra (DLOS)
measurements.
ODLTS measurements performed on oxygen-annealed AlGaN/
GaN(Fe) structures in Ref. [187] (Fig. 48) revealed the presence of
several hole traps H1 (1.1 eV, sp = 4.2  1012 cm2), H2 (0.41 eV,
2.3  1015 cm2), H3 (0.28 eV, 2.3  1016 cm2) not present in the
semi-insulating GaN(Fe) buffer probed by photoinduced current
transient spectroscopy (PICTS) spectra measurements (sp here is
the hole capture cross section). These hole-trap spectra are

Fig. 46. Band diagrams of AlGaN/semi-insulating GaN structures in accumulation


and depletion.
(After Ref. [237], Fig. 8.5a and b) Reprinted with permission.

ODLTS signal C/C (%)

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

30

E2(0.6 eV)

-10
-20 H3(0.28 eV)
-30
H2(0.41 eV)
E3(0.7 eV)
-40 E4(0.33 eV)
E1(0.6 eV)
H1(1.1 eV)
-50
-60
100 150 200 250 300 350 400
Temperature (K)

Fig. 48. ODLTS spectra measured on AlGaN/GaN(Fe) structures.

superimposed on the electron traps peaks of the opposite sign:


E1 (0.6 eV, 6.8  1015 cm2), E2 (0.59 eV, 3.3  1013 cm2), E3
(0.7 eV, 3.3  1015 cm2), E4 (0.33 eV, 1.3  1015 cm2). The E1 and
E2 electron traps are the ones that dominate the DLTS spectra in
Fig. 45.
Similar hole traps near Ev+1 eV and Ev+0.3 eV were previously
observed by PICTS in AlGaN/GaN heterostructures on sapphire and
on SiC. The magnitude of the hole traps peaks was much higher for
growth on sapphire [266]. The 1 eV hole traps are commonly
attributed to VGa complexes with shallow donors or to CN deep
acceptors (see the section on defects studies in GaN above). The
shallower hole traps were attributed in Ref. [266] to the hole traps
states on dislocations, possibly VGa-related, based on the results of
Ref. [230].
DLOS spectra were measured in some detail in Refs.
[237,244,274]. The authors of Ref. [274] performed comparative
measurements of DLOS spectra on AlGaN/GaN Schottky diodes in
partial depletion with the spectra on n-GaN samples. The results
are presented in Fig. 49. In GaN one can see three electron traps: T1
(optical ionization energy of 1.4 eV), T2 (2.64 eV), and T3 (2.9 eV).
These levels energies coincide reasonably with what is known
about deep traps in n-GaN. For the AlGaN/GaN structure one can
see in addition to the transitions occurring in the GaN buffer two
more traps, G1 (1.7 eV) and G2 (2.08 eV), corresponding to electron
emission from deep traps in the AlGaN barrier. There also exists a
hole-trap-like emission with optical threshold near 0.8 eV that
could correspond to recharging of hole traps in the lower half of the
AlGaN bandgap that are partly or fully ionized by the mechanism
outlined above. The traps in the AlGaN barrier, when photoionized,
could contribute to the 2DEG charge, but the overall density of
these traps determined from the change of the pinch-off voltage
and the change of capacitance under illumination was not high,
close to 2.2  1011 cm2.
DLOS spectra measured on AlGaN lms [244] revealed the
presence of acceptor states near Ev+1.2 eV and rather shallow

DLTS signal (Arb. units)

20
0.6 eV

15
10

O2 annealed
-3V->0V

5
0
-5

Reference

-10
-15
-20

0.1 eV

-3.2V->0V

100 150 200 250 300 350 400


Temperature (K)

Fig. 47. DLTS spectra measured near depletion for AlGaN/GaN structures before (the
virgin spectrum, black line) and after annealing in dry oxygen (blue line). (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)

Fig. 49. DLOS spectra of AlGaN/GaN and n-GaN.


(After Ref. [274], Fig. 3) Reprinted with permission, Copyright 2008 The Japan
Society of Applied Physics.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Capacitance (pF)

states near Ev+(0.150.2) eV (see the previous section). Similarly,


DLOS spectra measurements on AlGaN/GaN heterojunctions
indicated the existence in the AlGaN barrier of deep hole traps
near Ev+0.8 eV and Ev+0.150.2 eV [237].
Thus, the hole-traps-like signals detected in some DLTS spectra
measurements could indeed be related to the actual hole traps in
the AlGaN barrier if one assumes that such traps can be ionized via
tunneling in high electric eld. Such deep traps can contribute to
persistent photocapacitance effects for high photon energies, they
can also explain the hole-trap-like transients that are relatively
often observed in current DLTS measurements on HEMTs (see next
section). However, these traps are less detrimental to the device
performance than the capture and emission by electron traps.
Because DLTS measurements on AlGaN/GaN heterojunctions
are often handicapped by the series resistance effects they can, to
some extent, be supplemented by admittance spectra measurements with appropriately chosen biases and probing signal
frequencies. The model for the simple case of the absence of
strong metastabilities in CV characteristics has been developed in
Ref. [275]. This approach was rather successfully used by us to
compare the deep traps observed in AlGaN/GaN heterostructures
on Si substrate with the actual DLTS measurements on high-power
transistors with long multi-nger gates [267]. In Ref. [276] we
studied admittance spectra of AlGaN/AlN/GaN heterostructures
and observed transitions that could be ascribed to excitations from
the ground state to the excited state in the AlN/GaN quantum well.
Also, for high Al mole fraction heterostructures we reported the
appearance of additional interfacial defect states with activation
energy near 0.2 eV and 0.25 eV. The natural limitations here are
related to the general limitations of admittance spectroscopy
where one can only probe the traps near the Fermi level.
In Ref. [276] we also reported an interesting phenomenon in
persistent admittance spectra: the capacitance/conductance spectra showed steps/peaks with the frequency dependence opposite
to the standard admittance spectroscopy dependence (i.e. the peak
temperature shifted to lower temperature with increased probing
frequency). The activation energies determined from such
measurements were close to the activation energies observed
for the same samples in reverse DLTS experiments (see e.g. Fig. 50).
The explanation we offer for the effect is that, at high reverse
voltage, the dark CV characteristics are determined by the series
resistance effects caused by undercutting the 2DEG conductivity
below the Schottky diode, as in Fig. 42. Illumination at low
temperature persistently shifts the threshold voltage to more
positive values and decreases the series resistance effects in
capacitance and conductance versus frequency dependences. As
the sample freezes off the capacitance/conductance returns to the
dark conditions, the sooner the higher the probing frequency. This
approach could be used to estimate the thermal ionization energy
of the traps responsible for metastability.

1750

50% Al, -2.5V, PPC

1250
750

10 kHz

250
600

G/ (pF)

0.2 kHz

400

0.9 eV
10 kHz
0.2 kHz

200
0
50 100 150 200 250 300 350 400 450
Temperature (K)

Fig. 50. Admittance spectra measured on AlGaN/AlN/GaN structure after


illumination at low temperature.

31

We summarized in Table 3 the energy levels and capture cross


sections of deep traps detected by DLTS, ODLTS, DLOS and
admittance spectroscopy in AlGaN/GaN structures together with
the assumed location of deep traps. Comparison with traps
measurements on transistors in the next section allows to nd
many similarities thus justifying the effort, but also suggesting
more systematic studies.
5.2. Trapping in transistors
GaN-based eld effect transistors (FETs) have demonstrated
outstanding performance in terms of power density, operation
frequencies, breakdown voltages, operation temperatures (see e.g.
a recent review in Ref. [277] and references therein). However,
trapping by defects in such devices is still very much an issue and is
the main obstacle to much wider practical use of GaN transistors.
Trapping manifests itself in multiple phenomena, such as lower
output power at high frequencies compared to that expected from
DC characteristics, frequency dispersion, noise in devices, gate lag
and drain lag, high leakage currents, low breakdown voltage,
device degradation under operation or after irradiation, high
subthreshold currents (see e.g. Refs. [278288] to name a few). The
main effects relevant to the physics underlying all these processes
are the gate lag and drain lag. The early experiments dened the
major factors inuencing these processes and helped to develop
the models explaining experimental ndings and serving to
mitigate the negative inuence of trapping.
The gate lag basically consists in the appearance of more or less
long time delays of the drain current following the switching of the
gate voltage. The effect is clearly caused by tunneling of electrons
from the gate at high reverse voltage and capture of these electrons
by the states in the AlGaN barrier. This has been directly
demonstrated by scanning Kelvin probe measurements in Ref.
[263] that showed simultaneous build-up of the surface potential
near the drain edge of the gate and decrease of the drain current
upon application of the reverse bias pulse to the gate. After the gate
voltage was returned to 0 V the drain current gradually increased
simultaneously with changing the surface potential (see Fig. 51).
The spatial extent of the changed surface potential was some
tenths of a micron (Fig. 52). Illumination decreased the magnitude
of the changes in the surface potential and drain current [263]. The
effect was explained by increasing the negative charge on the traps
in the vicinity of the gate edge, Qt. This charge causes respective
decrease of the 2DEG density under the affected region Dq2DEG
according to relation [262]:

Dq2DEG Q t 1  d2t =d;

(7)

where d2t is the distance of the trapped charge from the 2DEG
region, d is the barrier thickness (for the charge trapped at the
barrier the change in the local 2DEG density is equal to the trapped
charge). It was assumed that the excessive charge is trapped by the
states on the surface of the AlGaN barrier forming the so called
virtual gate, i.e. effectively increasing the gate area and
decreasing the drain current in depletion [289]. The drain current
relaxation then reects the emission process from the traps in the
virtual gate region of AlGaN. Illumination is believed to create
electronhole pairs so that holes travel to the surface and
recombine with the trapped electrons while the photogenerated
electrons replenish the lost 2DEG charge.
The characteristics of the traps can be determined by analyzing
the magnitude of the drain current transients and the temperature
dependence of the recovery time. The former provides the estimate
of the density of the traps. The latter gives the value of the emission
rate of the trap in question [262]. The recovery process is usually
the sum of several exponents corresponding to different traps. The
relative contribution of each trap can be regulated by changing the

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

32

Table 3
Trap parameters detected from deep traps studies on AlGaN/GaN heterostructures.
Trap name

Type (e, n)

Ea (eV), Ec  Ea for e-traps;


Ev + Ea for h-traps

E1
E2
E3
E4
H1
H2
H3
E1a
H1a
A1
A2
A3
Ax
H1b
H2b
A
B
C
D
E
F
E1c
Hx
ETB1
ETB2
ETB3
ETS1
ETS2
ETS3

e
e
e
e
p
p
p
e
P?
e
e
e
e
p
p
e
e
e
e
e
e
e
p
e
e
e
e
e
e

0.6
0.59
0.7
0.33
1.1
0.41
0.28
0.6
0.82
1
1.2
1.3
0.9
1.24
1.3
0.15
0.21
0.12
0.42
0.49
0.94
0.3
0.82
0.15
0.29
0.4
0.18
0.27
0.45

s (cm2)

Comments
6.8  1015
3.3  1013

15

3  10

1.3  1015
4.2  1012
2.3  1015
2.3  1016
1.8  1016
7.8  1015
2  1012

5  1012
8.9  1019
1.1  1018
1.8  1018
1.3  1019
2.4  1014
1.1  1022d
3.6  1019
2.9  1014
5  1018
1.6  1013
6  1015
3  1015
4.9  1013
3  1013

[186,187], DLTS, ODLTS, buffer/interface


[186,187], DLTS, ODLTS, buffer/interface
[186,187], DLTS, ODLTS, buffer/interface
[186,187], DLTS, ODLTS, buffer/interface
[186,187], ODLTS, barrier
[186,187], ODLTS, barrier
[186,187], ODLTS, barrier
[25,32], PICTS, buffer
[25,32], PICTS, buffer
[221,222], DLTS, buffer, defect on dislocation
[221,222], DLTS, buffer, Ni defect on dislocation?
[222], DLTS, buffer, defect on dislocation, high C in the buffer
[222], DLTS, buffer, defect on dislocation, low C in the buffer
[222], DLTS, barrier, low C in the buffer
[222], DLTS, barrier, high C in the buffer
[271], DLTS, buffer/interface
[271], DLTS, buffer/interface
[271], DLTS
[271], DLTS
[271], DLTS
[271], DLTS
[272], DLTS, buffer
[273], DLTS
[267], Admittance, barrier
[267], Admittance, barrier
[267], Admittance, barrier
[267], Admittance, buffer
[267], Admittance, buffer
[267], Admittance, buffer

Traps notation not done in the paper.


H1 and H2 in the origin.
c
E1 in the origin.
d
There must be a typing mistake in the original paper, the peak with such level and such capture cross section cannot be observed at such temperature at the given
emission rate, most likely, should be 1.1  1012 cm2 as for the A1 trap.
b

amplitude of the negative gate bias pulse and thus the number of
electrons tunneling into AlGaN [262]. Then by measuring the
temperature dependence of recovery time for each component one
can nd the activation energy of emission for all traps. By varying
the drain voltage one can also study the effect of electric eld on
the measured activation energies [262].
Experimental measurements show that the eld dependence is
the one expected for the PooleFrenkel mechanism [262]. Thus, the
correction to the measured activation energy of the trap emission
is DfPF = (q3/pe)F1/2 and the activation energy of the trap Et

Fig. 51. The time evolution of the drain current and the surface potential measured
by the Kelvin probe for a HEMT structure after the application of the lling pulse
VG = 12, VD = 20 V.
(After Ref. [263], Fig. 3) Reprinted with permission, Copyright IEEE, 2003.

measured from the temperature dependence of the emission rate


en is given by Et = Et(0)  DfPF (q here is electronic charge, e the
dielectric permittivity, F the electric eld, Et(0) is the zero-eld Et
value).
Measurements of the on-resistance of transistors ROn dependence on the gate length for the xed sourcedrain distance show
that tunneling mainly occurs near the drain edge of the gate
because of the electric eld crowding in this region [290]. Characteristic relaxation times measured for different transistor struc-

Fig. 52. Spatial variation of the surface potential with time after the lling pulse.
(After Ref. [263], Fig. 4) Reprinted with permission, Copyright IEEE, 2003.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

tures widely vary from microseconds to tens of milliseconds. The


centers reported showed the apparent activation energies 0.22 eV
[262,291], 0.11 eV [262,286], 0.14 eV [262]. However, one has to
bear in mind that the eld in the transistors can be quite high
causing a large PooleFrenkel correction to the Et(0) energy, so
that, for example, the 0.11 eV trap detected in Ref. [262] had the
low-eld activation energy of 0.39 eV, while the 0.14 eV signature
belonged to the trap with Et(0) = 0.54  0.05 eV.
From the picture outlined above it becomes obvious that in
order to alleviate that kind of gate lag one has to decrease the
electric eld at the drain edge of the gate, to suppress as much as
possible the density of traps participating in the effect and to
minimize the tunneling component in reverse current. The electric
eld crowding near the edge can be seriously suppressed by
application of eld plates which has been indeed demonstrated in
a number of papers (see e.g. Refs. [291,292] among many others).
The same effect can be achieved by surface passivation by
dielectric layers, such as SiNx, SiO2, Sc2O3, particularly in
conjunction with eld plating (see e.g. Refs. [262,289,293297]).
Dielectric layers with high dielectric constant as, for example,
HfO2, look very promising in that respect [298]. The effect of
passivation on deep traps spectra performed by several groups also
suggests that passivation can decrease the density of dominant
surface traps. The traps spectra in transistors can be studied by
various versions of DLTS. In Ref. [299] the authors proposed a
version of current DLTS (CDLTS) in which the drainsource voltage
of HEMT is pulsed and the current DLTS spectra registered.
Measurements performed on unpassivated and passivated HEMTs
revealed the dominant role of traps with apparent activation
energy of 1.45 eV in both spectra, but a very strongly reduced
amplitude of the peak in the passivated device. The authors note
that the 1.45 eV trap is strongly reminding the trap responsible for
the 2DEG formation in the surface donor model discussed in the
previous subsection. In a later paper [300] the authors observed a
strong decrease of the magnitude of the peaks attributed to deep
traps with activation energies from 0.19 eV to 1.05 eV in SiNx
passivated HEMTs. The decrease of the dominant 1.05 eV trap was
particularly strong (the measurements were again done by drain
current DLTS with drain voltage pulsing). Current DLTS and PICTS
measurements on AlGaN/GaN HEMTs passivated by Sc2O3
deposition were performed by us in Ref. [301]. It was found that
the amplitude of the electron traps peaks belonging to Ec0.3 and
Ec1 eV levels were strongly suppressed by passivation. There are,
however, problems with these interpretations. If indeed the 2DEG
electrons are supplied by the dominant surface trap situated
around 1 eV below the conduction band of the AlGaN barrier, as
suggested in the s surface donor model, then passivation of such
traps should decrease the 2DEG density whereas it is well
established that in passivated HEMT structures the 2DEG density,
in fact, increases (see e.g. Refs. [263,289])? Also, the peak position
of the 1.45 eV trap near 300 K in Ref. [299] would require an
unrealistically high electron capture cross section for the trap.
Passivation of the Ec1 eV traps reported in Refs. [300,301] looks
more feasible, but then the main trap contributing to the surface
donor driven 2DEG formation could have been intact after the
passivation? The explanation of passivation effects offered in Ref.
[289] looks more plausible to us. The authors of [289] suggest that
surface passivation simply prevents the electrons tunneled into the
surface to be trapped by the surface states. Another possibility is
the build-up of the xed charge in the passivating dielectric
increasing the 2DEG density and stabilizing the surface upon
charge trapping [263].
Naturally, as the cause of the gate lag effects of the described
type seems to be due to high reverse leakage current causing
tunneling into defect sites and transport of the trapped charge
along the surface, decreasing the reverse current should generally

33

benet the HEMTs performance. As discussed in the previous


subsection the reverse current of Schottky diodes on AlGaN/GaN is
generally high and is to a large extent due to resonant tunneling via
defect states. It is strongly suspected that defect states on
dislocations play important part in these processes. We have
discussed theoretical considerations on that matter and briey
analyzed experimental results above. However, here it seems
appropriate to discuss the phenomena in relation to the surface
trapping in HEMTs in some more detail. Studies by conductive
atomic force microscopy and Kelvin probe microscopy of dislocations impact on carrier transport and trapping in n-GaN performed
in Ref. [63] showed that threading edge and mixed dislocations do
not contribute signicantly to leakage along the dislocation line,
but such dislocations are surrounded by an area of negative space
charge. The leakage along the threading dislocation lines is mostly
caused by threading screw dislocations that, however, do not seem
to be surrounded by space charge. The screw dislocations mostly
contribute to vertical leakage when they are open-core dislocations. This has been attributed to shallow donors (oxygen)
accumulation on the [10] facets of the open core screw dislocation
or of a nano/micropipe. That causes a local decrease of the width of
the space charge region in the Schottky diode [67,68]. When
comparing the leakage current of InAlN/GaN and AlGaN/GaN
Schottky diodes it has been reported that, with similar dislocation
densities in the barrier, the leakage current for InAlN barriers is
two orders of magnitude higher than for their AlGaN counterparts
which was ascribed to In accumulation in the dislocation core
[62]. But all these observations are not directly relevant to the
trapping phenomena we are discussing. Vertical tunneling,
possibly dislocations-mediated, can contribute to trapping below
the Schottky gates of HEMTs, but not necessarily in the virtual
gate region. What we are looking for are mechanisms enhancing
lateral tunneling into the bulk or the surface of the structure in
order for the virtual gate to be formed. This lateral tunneling can
be facilitated by threading edge dislocations because they are
surrounded by space charge and, in high electric eld, tunneling
might be possible into unoccupied states in the space charge
region. The trapped change can then move along the surface
between the dislocation sites via hopping or thermal activation.
For screw dislocations, particularly open core dislocations
surrounded by donors, tunneling can be enhanced by the lowered
local barrier thickness. The electrons tunneling into such dislocations can then move along the surface from dislocation to
dislocation till they are captured by a surface trap.
The mechanism of surface conductivity in AlGaN/GaN structures in electric elds typical of HEMT operation was studied in Ref.
[302] by a clever experiment exploiting an arrangement with the
ohmic source and two Schottky diodes depicted in Fig. 53. The
surface conductivity between the two Schottky diodes (Isurf) could
be separated from the vertical tunneling bringing electrons into
the 2DEG region from where they could be picked up by the ohmic
contact (Ib). The surface conductivity at relatively low electric
elds was found to be of Mott hopping type, with hopping,
presumably, occurring between the dislocation sites. At high
electric eld the thermally activated conductivity started to prevail
(Fig. 54). The mechanism of that conductivity was PooleFrenkel,
with the activation energy depending on the electric eld. This
conductivity was attributed to the above-barrier transport
between the dislocation sites. Naturally, for that type of a process,
decreasing the dislocation density should suppress both tunneling
into the surface states and transport of the trapped charge along
the surface, thus decreasing the amount of trapped charge and the
spatial extent of the virtual gate. Therefore, decreasing the
dislocation density by any means should be benecial to the
suppression of the gate lag effect. This could be achieved by growth
on more lattice matched substrates (GaN or AlN) or by carefully

34

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 53. Schematics of the surface conductivity measurements in AlGaN/GaN test structure imitating HEMT, (a) without SiN passivation of the surface between two Schottky
gates, (b) with SiN passivation.
(After Ref. [302], Fig. 1) Copyright AIP Publishing LLC, 2014.

optimizing the growth conditions (for example, in MBE growth it


was found that the density of leakage sites could be drastically
decreased by choosing the Ga/N ows ratio close to unity [303]).
Surface passivation by SiN decreasing the electric eld was found
to seriously decrease both the barrier leakage and the surface
conductivity as can be seen from Fig. 54. Annealing in oxygen
reducing the effects of surface leakage caused by residual surface
damage due to dry etching was also found to be rather effective
[186].
From what has been said it follows that substituting HEMT
structures with Schottky gates by MISHEMT structures with the
gate isolated with a dielectric layer can strongly suppress the gate
leakage and hence the gate lag, although trapping into states in the
dielectric layer has to be minimized. The gate leakage is indeed
decreased by such a procedure, but this is a special topic beyond
the scope of the present review.
In experimental studies of the gate lag effect various versions of
pulsing techniques were proposed, usually allowing to discriminate between trapping below the Schottky gate and in the gatedrain region. This is achieved by either applying gate pulses
without application of drain-current voltage or with a high draincurrent voltage providing a strong electric eld facilitating the

Fig. 54. The voltage dependence of the Schottky barrier (Ib) and surface conductivity
(Isurf) in the AlGaN/GaN test structure with and without SiN passivation.
(After Ref. [302], Fig. 2), Copyright AIP Publishing LLC, 2014.

formation of the virtual gate. The problem in such measurements is


often with the presence of several exponential components in the
charge build-up and charge recovery of the drain current. Joh and
del Alamo [304] proposed for these cases a convenient approach
that allows to trace the behavior of various components separately.
The experimentally observed current build-up or decay curves I(t)
are tted by the sum of several exponential processes (the authors
proposed to use 100 components) with respective weighing
P
amplitudes as in expression I(t) = aiexp(t/ti) + I1. The experiment I(t) curve then can be represented by the relaxation time
spectrum with several peaks reecting major processes.
The authors applied the procedure to the analysis of relaxation
curves in the AlGaN/GaN transistor structures. Fig. 55(a) gives an
example of the charge build-up process during the lling pulse of
the gatesource voltage VGS = 5 V for the device in the ON-state
(VGS = 1 V) and drainsource voltage VDS changing from 2 to 8 V, i.e.
in the linear ON-regime of the tested transistor. Fig. 55(b) shows
the result of the tting that demonstrates the presence of two main
processes TP1 and TP2. Variation of the VDS value demonstrates
that the characteristic trapping time for the process TP1 becomes
progressively shorter with increasing the drainsource DS voltage
VDS, whereas the characteristic trapping time of the TP2 process is
not affected (Fig. 56).
Measurements at different temperatures (Fig. 57) lead to strong
changes in the TP1 trapping time, but do not change the TP2 time.
The authors then applied the same procedure to the detrapping
current transients occurring upon application of a high gate
voltage of 12 V and a low drainsource voltage of 0 V. They
observed a single detrapping process DP1that was the reverse of
the trapping process TP1 and showed temperature activation of the
time constant with the activation energy of 0.57 eV. The DP2
detrapping process could only be observed together with the DP1
process after pulsing into the ON state with high drainsource
voltage. This process was not temperature activated. Based on
these measurements the TP1/DP1 process was assigned to
electrons capture/emission in AlGaN barrier (however, this
assignation is challenged by later results).
A similar approach to the relaxation time analysis of detrapping
processes in AlGaN/GaN MISHEMT structures on Si substrate was
applied in Ref. [305]. The devices were subjected to trap-lling
pulses in the off-state (8 V) and semi-on state (3 V) and drain
source voltages increasing from 0 to 200 V. The de-trapping

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

35

Fig. 57. The variation of the time constants of the TP1 andTP2 processes with
temperature.
(After Ref. [304], Fig. 5) Copyright IEEE, 2010.

Fig. 55. The current transient in AlGaN/GaN during trapping (a) and respective time
constants of the current transient TP1 and TP2.
(After Ref. [304], Fig. 3) Copyright IEEE, 2010.

Fig. 58. (a) RON time dependence and (b) Detrapping times spectra of AlGaN/GaN
MISHEMT on Si for pulsing into off-state and semi-off-state; two detrapping
processes E1 and E2 can be observed.
(After Ref. [305], Fig. 3a and b) Copyright 2014 AIP LLC.

Fig. 56. The evolution of the TP1 and TP2 processes time constants on the VDS value.
(After Ref. [304], Fig. 4) Copyright IEEE, 2010.

process was analyzed at VGS = 0 and VDS = 1.5 V. Two distinct


detrapping processes E1 and E2 were detected as shown in
Fig. 58. The off-state pulsing produced only the E2 state, the semion state pulsing gave rise to both the E1 and E2 processes. The
activation energies and electron capture cross sections deduced
from recovery measurements at different temperatures gave the
values of 0.6 eV, 1.2  1017 cm2 (E1) and 0.96 eV, 5.8  1015 cm2

(E2). The E1 trap was attributed to trapping in the AlGaN barrier or


in the GaN buffer. The E2 trap was assigned to trapping in the gate
edge/drain region. For AlGaN/GaN HEMTs with a thin low-Al AlGaN
back barrier off-state pulsing revealed a prominent role of electron
trapping in the back barrier that gave rise to the virtual gate
formation below the channel in the drain access region [306]. The
temperature measurements of detrapping again point to electron
traps near Ec0.6 eV as the main culprits.
A number of drain current DLTS spectra measurements
performed on AlGaN/GaN HEMTs also underlined the prominent
role of the Ec(0.570.6) eV traps [307309]. In the constant drain
current DLTS CIDDLTS two measurement regimes can be
employed [310]. In the rst case the device is pulsed from

36

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

pinch-off to transistor regime, the drain voltage is dynamically


adjusted to support the constant drain current ID and the change in
the drain voltage DVD with time is processed by the standard
gating technique of DLTS. The relaxation of the drain resistance
DRD = DVD/IDS (IDS = const) is monitored and produces in the
temperature scan a peak corresponding to electron trap with
activation energy Ea and capture cross section s. The concentration
of the traps nT can be calculated from:
nT n2s ns  Ldep:max =qmn DRd ;

(8)

where mn is the electron mobility, Ldep.max is the maximum extent


of the virtual gate charge region that can be either measured or
estimated from modeling as in Ref. [311]. The traps probed that
way are located in the drain access region.
In the gate-control CID-DLTS the drainsource voltage is low,
the lling negative gate pulse is regulated to sustain constant drain
current. The concentration of the traps is calculated from the
magnitude of the threshold voltage transient DVT as:
nT  2ebarrier DV T =qdbarrier :

(9)

In this version of the current DLTS technique the traps located


under the Schottky gate are predominantly probed. (It seems to be
implicitly assumed in Eq. (9) that the re-charged traps are located
at the surface; if one deals with traps that are e.g. uniformly
distributed along the length of the barrier respective correction
should be introduced.)
Comparative measurements of the drainvoltage CID-DLTS on
AlGaN/GaN and InAlN/GaN HEMTs in Ref. [309] showed the
predominance in both types of structures of the same electron
traps at Ec0.57 eV with s = 3  1015 cm2. The areal density of the
traps was quite high, 7  1012 cm2 in AlGaN/GaN structures and
8  1012 cm2 in InAlN/GaN structures (the Ldep.max in Eq. (8) was
determined by modeling to be 80 nm (AlGaN) and 72 nm (InAlN)
which was conrmed by the actual Kelvin probe measurements.
The traps were found to be responsible for drain lag (i.e. the
temporal lag of the drain current upon pulsing the drainsource
voltage) in both types of devices, the AlGaN/GaN and the InAlN
HEMTs (see Fig. 59(a) and (b)). From these observations it was

concluded that the traps in question should be situated in the GaN


buffer. Otherwise, for a barrier-related trap it would be difcult to
account for the same signatures of the traps in two seriously
different barrier types. However, as discussed in the previous
subsection, electron traps with the same signature were observed
in the AlGaN/GaN:Fe heterostructures with large area Schottky
diodes and were the main traps pinning the Fermi level in the
semi-insulating GaN:Fe buffer at Ec0.6 eV [186,187]. The density
of these traps in the buffer could not exceed 1016 cm3 which
would not produce the very high sheet densities of the traps
detected in Ref. [309]. It was assumed in [187] that these traps are
located in the GaN buffer close to the interface where the traps of
that kind pile up. It was observed that corresponding peak
magnitude increased substantially upon annealing at 750 8C in
oxygen. This suggests that the traps pile up can be facilitated by
high temperature treatments which would explain very high sheet
concentrations measured in the transistor structures in Ref. [309].
In Ref. [308] the Ec0.57 eV traps were also dominant in
passivated AlGaN/GaN HEMTs on SiC substrates. After 1000 h of
accelerated lifetime testing (ALT) at 260 8C and at high operating
current these devices showed measurable parameters degradation
manifesting itself in several times increased drain resistance. The
concentration of the Ec0.57 eV trap increased from
4.6  1012 cm2 to 7.3  1012 cm2 (the LDep.max in (2) was
measured to be 0.15 mm, in agreement with later measurements
of surface voltage spatial and time variations in Ref. [312]). This
increase in the main trap density was to the large part responsible
for the increase in gate lag observed after the ALT electric stressing
test. Similarly, a strong increase of the Ec0.6 eV traps was
observed after degradation of the AlGaN/GaN HEMTs in Ref. [313].
It is interesting, of course, to try to understand the origin of the
Ec0.57 eV traps. These traps should be omnipresent and thus
related either to a common technological impurity, such as C, or to
a structural defect, such as dislocations. As shown above, these
traps are located in the GaN buffer near the AlGaN/GaN interface.
In GaN, several traps with energies close to 0.6 eV have been
reported (see the Deep traps in GaN section above). The closest t
seems to be found with the deep trap near Ec0.6 eV associated

Fig. 59. (a) gate controlled CID-DLTS spectra in AlGaN/GaN and InAlN/GaN HEMTs showing the electron-trap-like state with level Ec0.69 eV in the AlGaN barrier and the
hole-trap-like state near Ev+1 eV in the InAlN barrier; (b) Arrhenius plots for respective traps.
(After Ref. [309], Fig. 4a and b) Copyright 2013 AIP Publishing LLC.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 60. The dependence on MOCVD growth temperature of the overall buffer traps
densities (solid circles) of the trap 1 density (open circles), trap 2 density (open
squares), and the sheet 2DEG charge density (solid diamonds) in AlGaN/GaN HEMTs
with semi-insulating GaN(C) buffers.
(After Ref. [318], Fig. 1) Copyright IEEE 2001.

1.6
DLTS signal (Arb. units)

with dislocations (the attribution is done based on the strength of


comparisons between DLTS spectra of polar and non-polar
undoped GaN lms [185] and measurements on ELOG GaN lms
with different thickness [184]). Detailed studies of the degradation
mechanism in AlGaN/GaN HEMTs (see e.g. the recent paper [314])
allowed to pinpoint the degradation to sites near the edge of the
gate giving rise to strong electroluminescence before degradation
and associated with pits after degradation. The fast degradation/
enhanced leakage sites were associated with dislocations, but the
real understanding of the processes that underline the effect still
has to be developed.
Other deep traps in AlGaN/GaN and InAlN HEMTs have been
examined in several papers. In Ref. [309] electron-trap-like defects
near Ec0.69 eV and hole-trap-like states near Ev+1 eV were
detected by gate-voltage pulsing CID-DLTS measurements respectively on AlGaN/GaN and InAlN/GaN HEMTs and were attributed to
traps in the barrier layer of the HEMTs (see Fig. 59). The density of
these traps was relatively low, about (0.31.6)  1010 cm2
[309]. The Ec0.69 eV barrier trap is similar to the traps detected
in the AlGaN barrier ODLTS spectra of AlGaN/GaN heterojunctions
(see Table 3 in the previous section and Ref. [187]). It is also similar
to the traps pinning the Fermi level in undoped semi-insulating
AlGaN lms grown by MBE (see the Fe doping and the Defects in
AlGaN sections above and Ref. [191]).
When the AlGaN/GaN transistor structures were prepared on
semi-insulating GaN buffers serious drain current lags resulting
from drainsource voltage pulsing were observed and were
attributed to electrons spilling from the channel at high drain
current and trapped by defects in the GaN buffer [285,315318].
(The effect is similar to backgating in GaAs-based transistors, see
the discussion in Ref. [315].) The trapping effects could be removed
by illumination of transistor structures due to the removal of the
trapped charge by light. Measurements of the spectral dependence
of the effect revealed the presence of two types of traps, Trap 1 with
the optical threshold of ionization near 1.8 eV and Trap 2 with the
optical threshold near 2.65 eV. This type of trapping was similar in
GaN MESFETs and AlGaN/GaN HEMTs proving the trapping in the
buffer to be the culprit. The effect was studied in detail for the
semi-insulating buffers doped with C and was attributed to the Crelated traps. This was conrmed by the traps density measurements on AlGaN/GaN:C HEMTs grown by MOCVD at various
growth pressures and hence various C concentration (see Fig. 60)
[318]. The traps density variations with growth pressure closely
followed the C concentration changes which proves the C-related
character of the Trap 1 and Trap 2 defects (Fig. 60). The drain lag of

37

ETB2

1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0

ETS2

-2V->1V

ETS1

-3V->-2.5V
ETB4
ETB3 ETS3

ETB1

ETB5

100 150 200 250 300 350 400


Temperature (K)

Fig. 61. Deep electron traps observed in one of the studied AlGaN/GaN/Si HEMTs by
DLTS; ETS traps with numbers are attributed to electron traps in the buffer and were
detected by applying the steady state bias of 3 V, with the voltage pulsed to
2.5 V, ETB traps are believed to be located mainly in the AlGaN barrier or at the
AlGaN/GaN interface and were detected with the 2V-> 1 V bias/pulse sequence.
The traps activation energies are shown below the traps label.

that type is not observed for HEMTs on conducting buffers


[285]. For other types of semi-insulating buffers (e.g. GaN(Fe)
buffers) the spectral dependences of the effect were not, to the
present authors knowledge, measured in any detail while such
measurements could be helpful to better understanding the
mechanism of the effect. For example, the Fermi level in GaN:C
semi-insulating lms is usually pinned near Ec0.8 eV [145]. Thus
the states with optical thresholds 1.8 eV and 2.65 eV should
normally be lled by electrons even in AlGaN/GaN HEMTs with a
strong band-bending in the buffer (see Fig. 46 in the previous
subsection). Then how does the trapping of electrons at high drain
currents occur?
As mentioned above standard capacitance DLTS measurements
are possible for high-power HEMT structures with long multinger gates having the large enough gate capacitance. Such
measurements were performed for AlGaN/GaN high-power HEMTs
grown on Si (see Refs. [265,269,270]). The traps in the buffer and in
the barrier could be separated by applying different steady-state
biases and varying the height of the injection pulse. Fig. 61
compares the traps observed in the buffer and in the AlGaN barrier
in one of the studied transistors (the traps labeled ETS refer to the
buffer traps, the traps labeled ETB are believed to be the AlGaN
barrier or interface traps). The variations of deep traps spectra
between different samples were rather strong, but the samples
could be roughly broken into two major groups. In group 1 the
shallow buffer and interface traps had high concentration, in group
2 only deeper barrier/interface traps were observed (Fig. 61
displays typical spectra of the group1 transistors, for one of the
group 2 transistors the barrier/interface spectrum is shown in
Fig. 62). The trap parameters detected in these measurements are
summarized in Table 4. Systematic comparison of transistor
characteristics and deep traps spectra has not been performed as
yet, but preliminarily one can say that the high concentration of
shallower ETB1, ETB2, ETS1, ETS2 traps in group 1 samples
increased the saturation drain current, but decreased the slope of
the drain-current IV characteristics in the linear region compared
to the group 1 transistors where only deeper ETB3, ETB4, ETB5
traps were detected.
The temperature dependence of capacitance and conductance
recovery time transients upon application of various gate voltages
was measured for wide comb-like Schottky diodes and compared
to the drain current relaxation times of 400 V AlGaN/GaN HEMT on
SiC [270]. The transistor characteristics were analyzed using the
relaxation times spectra characterization method of Ref.
[304]. The off-state 400 V trap-lling pulses at different temperatures close to room temperature were used for drain lag transistor
measurements. The test Schottky diodes and the transistor

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

38

DLTS signal (Arb. units)

3.0
2.5
2.0
0.8 eV
ETB5

1.5
1.0
0.5
0.0

0.55 eV
ETB4
100 150 200 250 300 350 400
Temperature (K)

Fig. 62. Deep traps spectra in AlGaN/GaN/Si HEMTs studied by DLTS.

structures were prepared on the same epitaxial wafer. Electrontrap-like states with activation energies 0.49 eV and acceptor-traplike states with activation energies 0.44 eV and 0.52 eV were found
in the drain access region of the Schottky diodes. For the barrier
region under the gate the hole-trap-like centers with activation
energy of 0.18 eV were dominant. The latter activation energy was
close to the activation energy obtained from the drain-lag
measurements on transistors (0.14 eV).
As discussed in the previous subsection, in our understanding,
the hole-trap-like features in transient measurements on HEMTs
should most likely come from tunneling from acceptors below the
Fermi level at high reverse voltages and subsequent hole emission
from these traps during recovery. The acceptors with energy close
to the hole traps observed in Ref. [270] have indeed been detected
in AlGaN barriers of AlGaN/GaN HEMTs (see the previous
subsection and Refs. [187,244]).
In terms of degradation mechanisms for high-off-state voltage
stressing several additional traps different from the above
discussed Ec0.57 eV trap have been observed and reported. The
early degradation stages in AlGaN/GaN HEMTs were reported to be
strongly affected by the dislocation density (Fig. 63). The amount
of degradation was low for low dislocation densities of
2  107 cm2, it was much higher for medium dislocation
density of 5  108 cm2, and the degradation became very
pronounced for dislocation densities over 1010 cm2. At that, the
recovery time spectra analysis revealed the presence of three
different recovery processes TP1, TP2, and TP3 [319]. The TP1
process became very dominant in samples with high dislocation

Table 4
Activation energies Ea and electron capture cross sections s of electron traps in the
barrier (traps labeled ETB with serial number) and in the buffer (traps labeled
ETS with serial number); the parameters were determined from DLTS
measurements with different biases and different injection pulse heights
performed on AlGaN/GaN/Si HEMTS with long multi-nger gates.
Trap signature

Ea (eV)

s (cm2)

ETB1
ETB2
ETB3
ETB4
ETB5
ETS1
ETS2
ETS3
ETS4

0.15
0.29
0.4
0.55
0.76
0.18
0.27
0.45
0.9

5  1018
1.6  1013
6  1015
5  1015
1.9  1014
3  1015
4.9  1013
3  1013
1  1013

density. For these samples the activation energy of the process was
found to be equal to 0.48 eV, as for one of the traps detected in Ref.
[270]. Previously similar traps have been detected in AlGaN/GaN
transistors after degradation and associated with movement of
oxygen from the surface oxide on AlGaN [320]. The authors of Ref.
[319] note that the dislocations, when present in high density, can
easily form reverse leakage channels near the edge of the Schottky
gate and can serve as the preferential sites for device breakdown
upon application of strong off-state stress. At the same time,
dislocations can produce avenues for easy electrically stimulated
diffusion of impurities. However, despite the enormous practical
importance, these issues need a lot more study.
Dislocations were found to also have a marked impact on other
transistor characteristics. In Ref. [321] the authors measured the
subthreshold drainsource current dependence on the draincurrent voltage for the pinched-off channel of AlGaN/GaN
transistor with semi-insulating buffer grown on semi-insulating
SiC. They observed the dependence that, when built in double
logarithmic scale, showed a clear evidence of the trap lling
current (TFC) regime [322]: the current rst increased linearly with
voltage, then grew superlinearly with voltage as I 1 Vn, and then
switched to the I 1 V2 region after a sharp current increase at the
breakdown voltage C. For the semi-insulating GaN single layers on
SiC the character of the current owing between the two contacts
on top was similar (Fig. 64). The switching voltage at point C was
found to increase quadratically with the layer thickness d and to

Fig. 63. Trapping in AlGaN/GaN HEMTs induced by OFF-state stress before and after degradation in transistors with different dislocation densities; the inset in the second
panel shows the temperature dependence of detrapping in process TP1.
(After Ref. [320], Fig. 3) Copyright American Institute of Physics, 2011.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 64. IV characteristic of one of the semi-insulating GaN samples showing


different regimes of trap lling current ow.
(After Ref. [321], Fig. 2) Copyright the Japan Society of Applied Physics, 2008.

39

Fig. 66. Drain current as a function of gate voltage at drain-source voltage of 2 V in


GaN(cap)/AlGaN/AlN/GaN transistors on sapphire.
(After Ref. [325], Fig. 3) Copyright IEEE 2003).

follow well the expression for the trap lling limit regime VTFL
[322]:
2

V TFL qN t d =2e;

(10)

where q is electronic charge, Nt is the trap density, and e is the


dielectric permittivity.
The calculated density of traps in the lms of different thickness
(Fig. 65) coincided very reasonably with the changes with
thickness of the dislocation density as determined by transmission
electron microscopy (TEM). Thus, dislocations seem to be
providing trapping sites for current ow through the semiinsulating GaN buffer and determine the breakdown voltage value
for the HEMTs in the OFF-state.
5.3. Radiation effects in GaN-based HEMTs
Radiation effects in AlGaN/GaN HEMTs have already been
described in recent reviews [17,18,323] and the interested reader
can nd the data on device parameters changes and references to
multiple original papers in these reviews. The interest to the topic
is justied by planned space and military applications of AlGaN/
GaN HEMTs. The conclusions of multiple research groups reported
so far are quite optimistic: the changes of AlGaN/GaN parameters
start at doses far exceeding the ones expected in space applications
and the radiation hardness of GaN-based devices is about two
orders of magnitude higher than for their Si and AlGaAs/GaAs
counterparts. Thus, there is no immediate practical need to
increase the radiation hardness of GaN HEMTs. Therefore, our
interest in this section is more a scientic one: we would like to
rationalize the observed parameters behavior in terms of the

Fig. 65. The dependence of the deep traps density responsible for the breakdown
voltage on the semi-insulating GaN thickness (and hence the dislocation density).
(After Ref. [321], Fig. 3) Copyright the Japan Society of Applied Physics, 2008.

Fig. 67. Threshold voltage variations with proton uence sapphire.


(After Ref. [325], Fig. 4) Copyright IEEE 2003).

impact of deep traps introduced by radiation in heterostructures.


The main bulk of experimental device results refers to proton
irradiation because of the practical importance of protons of (1
100) MeV for space applications [17,18,323] and because high
energy protons are convenient when studying the radiation effects
both on heterostructures and on ready devices. With neutron
irradiation problems with induced radioactivity are a serious
concern, while, for gamma-irradiation very high radiation doses
proved to be necessary to observe detectable changes. Electron
irradiation has been widely used for defect studies aimed at
determining fundamental material characteristics, but has not
been so popular in device degradation work [17,18,323].
With proton irradiation, the main effects in HEMTs were the
shift of the threshold voltage toward more positive values, the
decrease of the saturation current Idmax caused by the decrease of
the 2DEG density and 2DEG mobility, and the decrease of devices
transconductance gm (see Refs. [324328]). These effects are
illustrated by Figs. 6669 for 1.8 MeV proton irradiation at room
temperature. The rate of parameters changes was found to
decrease with increasing the proton energy which correlates with
the amount of energy deposited by the given particle into atomic
displacements in the AlGaN barrier and the 2DEG region [329
331]. The rate of positive threshold voltage shift upon 1.8 MeV
protons irradiation was found to be the highest for N-rich MBE
growth, considerably lower for Ga-rich MBE growth, and the
lowest for NH3-MBE and MOCVD growth (both characterized by

40

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 68. Transconductance variation with proton uence.


(After Ref. [325], Fig. 5) Copyright IEEE 2003).

Fig. 69. Saturation drain current variation with proton uence.


(After Ref. [325], Fig. 8) Copyright IEEE, 2003.

N-rich conditions and the abundance of hydrogen (see Ref. [229]


and Fig. 70 (bottom); Fig. 70 (top) shows the threshold voltage
changes in the same types of transistors upon application of the
OFF-stress during prolonged times).
For 60Co gamma-irradiation with a very high dose of 600 mrad,
the threshold voltage shifted to more negative values, the
saturation current slightly increased, while the slope of the
drainsource IV characteristics in the linear region markedly
decreased signifying a slightly higher 2DEG concentration, but a
lower 2DEG mobility (see Ref. [332] and Figs. 71 and 72).
Electron irradiation allows to generate much higher densities of
defects than the gamma-radiation. Under electron irradiation of
HEMT structures one observes negative-voltage threshold shifts at
lower uences similarly to the case of gamma-radiation, but
positive shifts for higher uences, similarly to the case of protons
(see Ref. [182] and Fig. 73; mind that, for NH3-MBE and MOCVD
HEMTs in Fig. 70 (bottom) the threshold voltage also initially, at
low proton uences, shifts to more negative values). The rate of
electron irradiation induced parameters changes of the AlN/GaN
HEMTs was found to be about an order of magnitude slower than
that of AlGaN/GaN HEMTs. This was explained by the much thinner
barrier in the AlN case and, hence, the lower energy impacted into
lattice defects formation by irradiating electrons in such a barrier
(see Ref. [182]).

Fig. 70. (Top) Threshold voltage evolution in AlGaN/GaN HEMTs after electrical
stress as a function of stress time; (Bottom) threshold voltage evolution after proton
irradiation for AlGaN/GaN HEMTs; the results are shown for structures grown by
MBE under N-rich and Ga-rich conditions, structures grown by NH3-MBE, and
MOCVD-grown structures.
(After Ref. [229], Fig. 2) Copyright IEEE, 2011.

Currently it seems to be widely accepted that the positive shifts


of the threshold voltage are caused by the introduction of
negatively charged traps in the AlGaN barrier or in the GaN buffer,
while negative shifts are due to the increase of the positive charge
in the barrier. The authors of Ref. [229] calculated the defects
formation energies and defects concentrations in AlGaN by
protons. They argue that the type of defects formed by irradiation
depends on the density and type of starting defects. If, for example,
in the initial state the dominant defects are VGa acceptors, as it
should be for n-type material, particularly under N-rich growth
conditions, proton irradiation can transform them into negatively
charged nitrogen interstitials Ni and negatively charged VGaVN
divacancies (the latter with the 2/3 charge transition CT level
near Ec1 eV in AlGaN [229]). The result is the build-up of the
negative charge in the barrier and the shift of the threshold voltage
to more positive values. At the same time, as the CT level of
divacancies is close to the Fermi level in the barrier, one can expect
the increase in the low-frequency noise of the devices which
indeed is observed [229]. Since the density of VGa should be higher
under the N-rich conditions in MBE the value of the positive shift in

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 71. Transconductance of AlGaN/GaN HEMTs before and after the 60Co gammairradiation.
(After Ref. [332], Fig. 2) Copyright American Institute of Physics, 2002.

the threshold voltage should be the more pronounced than for MBE
growth under Ga-rich conditions. This is conrmed by experiment
as illustrated by Fig. 70 (bottom). Apparently, for the MOCVD and
NH3-MBE transistor structures additional hydrogen passivation of
defects should be invoked to explain the relatively lower shifts of

Capacitance (pF)

Fig. 72. AlGaN/GaN transistor IVs before and after 60Co gamma-irradiation.
(After Ref. [332], Fig. 3) Copyright American Institute of Physics, 2002.

1000

AlGaN/GaN
AlN/GaN

100

10
-8

-6

-4

-2

Voltage (V)
Fig. 73. CV characteristics of AlGaN/GaN (solid lines) and AlN/GaN (dashed lines)
measured at 10 kHz before 10 MeV irradiation (black curves) and after irradiation
with the uences of 5  1015 cm2 (red curves), and 1016 cm2 (blue curves). (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)

41

the threshold voltage in Fig. 70 (bottom). This defects hydrogenation was proposed as the cause of the negative threshold voltage
shifts upon the application of OFF-stress, as in Fig. 70 (top)
[229]. Here the dominant hydrogenated defects are supposed to be
C acceptors and nitrogen antisite donors (see also Refs.
[229,333,334]).
The model does explain some important features of the
transistors degradation under electrical stress and after proton
bombardment. The rational point also is that, at low radiation
doses, the sign of the threshold voltage shift can be negative and
coincide with the sign of the threshold voltage shift during
electrical OFF-stress degradation, presumably, because at these
early irradiation stages radiation causes the same modication of
defects as the electric stress. This situation is indeed encountered
for proton, electron, and gamma-irradiations and could well be
related to dehydrogenation of structural defects or impurities at
low doses. However, further developments and renements of the
model are necessary. For example, from detailed studies of VGa
behavior in III-Nitrides it seems clear that in most cases they are
complexed with oxygen, the NGa antisites are believed by some
groups never to be the defects present in high concentration in asgrown III-Nitrides, simultaneous hydrogen passivation of C
acceptors and NGa antisites donors requires hydrogen to co-exist
in the donor and acceptor state which can only occur for an arrow
range of Fermi level positions (see the section on theory of defects
above). Besides, as weve seen in the previous subsection,
dislocations and impurities play an important role in device
degradation under stress and perhaps should not be discarded in
modeling of radiation damage in devices.
Finally, the results reported in Refs. [268,269] suggest that the
defects responsible for the positive shift of the threshold voltage
after irradiation are the same defects as present in the material
rather than new types of defects as suggested in Ref. [229]. Namely,
in Refs. [268,269] it was shown that neutron and electron
irradiations produce defects similar to the defects causing
metastabilities and low threshold voltages in AlGaN/GaN,
AlGaN/AlN/GaN, and InAlN heterostructures (see the subsection
on heterostructures above). Neutron irradiation effects were
studied in Ref. [268] for a group of AlGaN/AlN/GaN heterostructures with Al composition in the barrier of 20%, 30%, 40%, and 50%.
These were the same heterostructures for which high acceptor
concentrations in the AlGaN barrier were detected in CV
measurements (see above). The neutron irradiation increased
the density of these barrier acceptors and the acceptors in question
were shown to be the same that cause metastable decrease of the
threshold voltage upon cooling down at high reverse bias and
persistent increase of the threshold voltage after low temperature
illumination. The threshold voltages necessary for 2DEG depletion
in Ni Schottky diodes were progressively shifting toward more
positive values with increasing the neutron uence in the (1
5)  1015 cm2 range. The reverse DLTS and the optical ionization
spectra of acceptors before and after irradiation were very similar
to these characteristics in the starting lms. Thus, it was concluded
that the neutron irradiation increases the density of pre-existing
barrier traps rather than creates new traps. The traps concentrations at various neutron uences could be estimated from observed
changes of the threshold voltage and increased approximately
linearly with the neutron uence (see Fig. 74). The introduction
rate of these acceptors coincided very reasonably with the
introduction rate of compensating acceptors measured for
undoped AlGaN single lms in Refs. [247,335]. Interestingly and
somewhat counter-intuitively, the introduction rate of acceptors
increased for AlGaN lms with high Al composition (40% or higher)
for which the starting concentrations of deep barrier acceptor traps
were high. Most likely, this increased starting concentration
reected higher density of native defects and impurities (see

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

42
12

5x10

12

40% Al

-2

Nss (cm )

4x10

12

3x10

30% Al

12

2x10

20% Al

12

1x10

15

15

1x10
2x10
-2
Neutron fluence (cm )

3x10

15

Fig. 74. The areal density of deep acceptor traps in the barrier of AlGaN/AlN/GaN
heterostructures as inuenced by neutron irradiation, the curves are marked by the
Al composition in the AlGaN sub-barrier of these structures.

above). The lower radiation tolerance of these more defective and


less pure materials is, however, in line with the results of
calculations in Ref. [229] suggesting that the effective radiation
defects formation energies should decrease for material with a
higher starting density of defects.
The second effect of neutron irradiation reported in Ref. [268]
was the decrease of the 2DEG mobility measured by Hall effect.
This is illustrated by Fig. 75. The characteristic feature of these
results was that the mobility started to rapidly decrease at a certain
threshold uence, and the value of this uence was the lower the
higher the starting density of the barrier traps. Simultaneously
with the strong decrease of 2DEG mobility we observed measurable decrease of the Schottky diodes capacitance in accumulation.
For uences greatly exceeding the threshold uence the 2DEG
conductivity could no longer be observed. In the sample with the
highest Al composition of 50% in the AlGaN barrier and the highest
density of deep barrier acceptors this loss of 2DEG mobility
occurred after irradiation with the lowest neutron uence used
(1015 cm2). The mobility changes observed were explained in the
following way. The Fermi level at the surface of AlGaN/GaN
heterostructures is pinned somewhere around Ec1 eV. Thus, at
0 V surface potential the 2DEG density should be at its maximum
at the AlGaN/GaN interface. However, the presence of a high
density of deep charged acceptors in the barrier causes local
uctuations of the threshold voltage and hence nonuniform 2DEG
concentration and additional scattering that affect the 2DEG
mobility. The magnitude of the effect is higher for higher density of
the barrier acceptors. As the density of these acceptors increases
with irradiation the magnitude of the threshold voltage uctuations increases causing the gradual decrease of mobility. When the
magnitude of local uctuations of the threshold voltage becomes
so high that in parts of the structure the voltage locally exceeds the
threshold voltage, local areas where no 2DEG is present appear.

This results in the current ow in the 2DEG region acquiring the


percolation character thus pulling down the effective 2DEG
mobility. The Schottky diodes capacitance in accumulation also
decreases because only part of the area is occupied by the 2DEG
electrons. Such situation corresponds to the threshold neutron
uence case in Fig. 75.
The behavior of AlGaN/GaN heterostructures with Al composition of 30% was also compared with the behavior of the AlGaN/AlN/
GaN heterostructures for the case of neutron irradiation. The
presence of deep acceptors causing the same metastability as for
the AlGaN/AlN/GaN structures was observed and the characteristics of acceptors as determined by the spectral dependence of the
threshold voltage shift and the signatures of defects in reverse
DLTS were similar for AlGaN/AlN/GaN and AlGaN/GaN structures.
The behavior of mobility with irradiation and the changes of the
threshold voltage with neutron uence were also similar for the
same concentration of Al in the barrier indicating that the main
event occurred either in the AlGaN barrier or in the GaN buffer, not
at the AlN/GaN interface (see Fig. 75 for mobility changes; the
higher starting mobility in the AlGaN/AlN/GaN structure is to be
expected because of the suppressed alloy scattering by the AlN
inset (see e.g. Ref. [182])).
For InAlN/GaN heterostructures, the neutron irradiation effects
again were similar to those observed in AlGaN/AlN/GaN, AlGaN/
GaN heterostructures, but the density of deep compensating
acceptors before irradiation was quite high, on the level of AlGaN
barriers with high Al mole fraction (3.4  1012 cm2 in InAlN
versus 1.8  1012 cm2 for AlGaN (40% Al) and 3.2  1012 cm2 for
AlGaN (50% Al)). This could be related to the presence of higher
impurity (particularly, oxygen) concentration in InAlN (see the
section on donor centers in III-Nitrides above). The starting 2DEG
mobility value in InAlN/GaN heterostructures was quite comparable with the 2DEG mobility in AlGaN/GaN structures, but the
threshold neutron uence for the onset of strong mobility
degradation was much lower than for AlGaN (see Fig. 75).
Similar effects were observed for AlGaN(30% Al)/GaN/sapphire,
AlGaN(30% Al)/AlN/GaN/sapphire, AlGaN (30% Al)/GaN/Si, and
InAlN/GaN/sapphire structures grown by MOCVD either on
sapphire or Si substrates and irradiated at room temperature
with 10 MeV electrons with uences in the 2  1015
3  1016 cm2 range. Fig. 76 shows the 2DEG mobility evolution
with electron irradiation for the studied heterostructures. The
positive shifts of the threshold voltage of CV characteristics and
the metastable changes in the threshold voltage upon cooling
down and illumination had the same nature as for neutron
irradiation. The mobility degradation occurred in a similar fashion
for all the AlGaN heterostructures and was not strongly affected by
the presence of AlN inset layer or changing of the substrate from
sapphire to Si. But, as for the case of neutron irradiation, the
decrease of mobility with electron uence was the fastest for the

30% Al
1400
2

1200
AlGaN/GaN
1000
40%

800
600
InAlN/GaN
400

15

15

2x10
3x10
1x10
-2
Neutron fluence (cm )

2DEG mobility (cm /Vs)

1600

2DEG mobility (cm /Vs)

1600

15

Fig. 75. The 2DEG electron mobility variations with neutron uence for AlGaN/AlN/
GaN heterostuctures with Al concentration in the AlGaN barrier of 30% and 40%;
also shown are the data for the AlGaN/GaN heterostructures (30% Al) and for InAlN/
GaN heterostructures.

1400

AlGaN/AlN/GaN

1200
1000
800

AlGaN/GaN
InAlN/GaN
AlGaN/GaN/Si

600
0

16

16

1x10
2x10
3x10
-2
Electron fluence (cm )

16

4x10

16

Fig. 76. The 2DEG electron mobility evolution with 10 MeV electrons uence for
AlGaN/AlN/GaN, AlGaN/GaN, and InAlN/GaN heterostructures grown on sapphire,
also shown are the results for AlGaN/GaN/Si heterostructures grown on Si substrate.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Reference

10

80
DC
100 Hz
10 kHz
VDS=+5V

8
6
4

60
40
20

2
0
-4.0

-3.5

-3.0

-2.5

-2.0

VG (V)

-1.5

100

12

IDS (mA/mm)

Normalied IDS (%)

100

(a)

16
2
DC
1.6x10 /cm
100 Hz
10 kHz

10
8

80
60

VDS=+5V

40

4
(b)

2
0
-4.0

-3.5

20
-3.0

-2.5

-2.0

-1.5

Normalized IDS (%)

IDS (mA/mm)

12

43

VG (V)
Fig. 77. (a) DC and AC measurements of the drain current as a function of gate
voltage for the AlGaN/GaN/Si HEMT before electron irradiation; (b) the same after
irradiation with 1.3  1016 cm2 10 MeV electrons.

InAlN/GaN/sapphire heterostructure, obviously, for the same


reasons as discussed above for the case of neutrons (see Fig. 10).
We also performed some preliminary experiments on the
amount of trapping introduced in high-power AlGaN/GaN/Si
HEMTs by electron irradiation. Fig. 77(a) and (b) shows the
changes in the gate lag produced by irradiation with electron
uence of 1.3  1016 cm3. The difference between the drain
source current IDS measured as a function of gate voltage VG at DC
and AC conditions was taken as a measure of gate lag magnitude
(the measurements were done at the drainsource voltage of 5 V
and the pulsing frequency in AC of 100 Hz and 10 kHz, the
technique is described in detail in Ref. [336]). It can be seen that the
AC current in accumulation constituted about 60% of the DC
current before irradiation, but only about 30% after irradiation.
DLTS spectra measured on this device before and after irradiation
are compared in Fig. 78 (Ref. [269]). It can be seen that electron
irradiation introduced a high concentration of electron traps with
activation energy 0.17, 0.3, 0.45 eV not observed in this particular
transistor before irradiation. It also quite measurably increased the
density of electron traps with activation energy 0.55 eV and 0.8 eV
detected before irradiation. Temperature measurements of the
transient effects in transistor characteristics have not as yet been
performed for these transistors to nd which of the traps whose

DLTS signal (Arb. units)

3.0
2.5
2.0
0.3 eV

1.5

0.55 eV
0.45 eV

1.0
0.5
0.0

0.8 eV
16
2
1.3x10 e/cm

0.17 eV
before
100 150 200 250 300 350 400
Temperature (K)

Fig. 78. DLTS spectra in the virgin AlGaN/GaN/Si HEMT and the same HEMT after
irradiation with 1.3  1016 cm2 10 MeV electrons.

Fig. 79. (a) Drain-source current transients before and after proton irradiation, (b)
time-domain analysis of the data in (a).
(After Ref. [337], Fig. 4a and b) Copyright Elsevier Ltd., 2014.

density was enhanced by irradiation contributed most to the


radiation enhancement of the gate lag. Such measurements are
denitely necessary in order to understand the nature of changes
in transistor characteristics.
Some experiments of that sort are described in Ref. [337] for
proton irradiated transistors designed for use in space applications.
The authors found that 3 MeV proton irraditation of AlGaN/GaN
HEMTs prepared on SiC substrates did not vary much the DC
characteristics of devices for uences up to 1014 cm2. However,
after irradiation with the highest proton uence of 1014 cm2, a
very marked increase of the trapping impact on drain current was
observed with pulsed measurements. The time-domain analysis of
the transients produced two distinct peaks E2 and E4 corresponding to the centers responsible for the gate and drain-lag
phenomena (Fig. 79). The density of both traps increased after
proton irradiation. Measurement of the effect at various temperatures gave the activation energies and the capture cross sections of
the E2 and E4 electron traps as 0.6 eV, 3.9  1015 cm2 (E2) and
0.8 eV, 4  1015 cm2 (E4). The results are very similar to those we
observed in electron irradiation of AlGaN/GaN/Si HEMTs in Ref.
[269], and the E2 and E4 traps parameters in Ref. [337] are very
reasonably close to the ETB4 and ETB5 electron traps parameters
determined in DLTS measurements on AlGaN/GaN/Si HEMTs (see
Ref. [267] and Table 4 above). Both types of traps increase in
concentration under high-energy particles bombardment. The E2
and E4 traps are believed by the authors of Ref. [337] to be located
in the region below the gate which seems to be in agreement with
conclusions of DLTS-based measurements. The E2 (ETB4) traps role
in changing the trapping in AC transistor parameters and their
degradation with electrical stressing has already been discussed in

44

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

the previous subsection, where the traps in question were assigned


to defects in the GaN buffer at the AlGaN/GaN interface. But of
course, more work along these lines is necessary.
As argued above, the changes in the threshold voltage and in the
channel mobility of AlGaN/GaN transistors can be to some extent
explained by the introduction of deep acceptors in the barrier layer
of the structures. However, deep traps are also introduced in the
GaN buffer layer of HEMTs and can lead to similar effects. Thus,
modeling of proton damage in AlGaN/GaN HEMTs based on
assuming the introduction of deep traps contributing to partial
2DEG depletion and to 2DEG electrons scattering was successful in
predicting the devices performance with radiation (see Ref. [338]).
These studies are very important, particularly for understanding
the nature of current collapse in various AlGaN/GaN HEMTs
structures and more research is denitely necessary on that front.
Deep traps contribution to the AlGaN/GaN HEMTs degradation
after bombardment with high energy particles is very important,
but not exclusively important. Issues such as surface passivation,
Schottky diodes and ohmic contacts performance are also very
important. For example, it has been shown that SiN passivation of
the surface of the AlGaN/GaN HEMTs seriously improves the
tolerance to the 60Co gamma-irradiation (see e.g. Ref. [339]),
dielectric passivation was also observed to improve matters in
proton irradiated HEMTs, see e.g. Refs. [327,328,330]. Importantly,
it seems that, for the protons uences used, the defects in the
passivating dielectric were not causing new problems. Radiation
induced variations in Schottky diodes quality can seriously impact
the trapping phenomena via changing the reverse current leakage.
The general trend seems to be that low radiation doses adversely
affect the reverse current, while high doses decrease it. These
issues, though very important, are beyond the scope of the present
review. Interested reader can nd the recent results and some
relevant references in Ref. [340].
6. Deep traps in GaN-based LEDs and LDs
At rst glance studies of deep trap defects in GaN-based LEDs
and LDs seem to be a purely scientic exercise since, in most cases
of practical interest, one would be concerned with the devices
performance at high output power. Under these conditions the
contribution of SchockleyReadHall (SRH) non-radiative recombination channel in the active region is usually not particularly
signicant, the dominant cause limiting the performance being
Auger recombination and the spatial separation of electrons and
holes in the quantum well due to the impact of polarization eld
(see e.g. the discussion in Ref. [341]). This, however, becomes less
obvious for devices operation at elevated temperatures where SRH
recombination can again become an important factor, together
with the charge carriers spilling out of the active region (see e.g. the
discussion in Ref. [342]).
Moreover, for LEDs and lasers operating in the green-red
spectral region, the impact of deep traps can be signicantly higher
than for blue LEDs and LDs because of the increased concentration
of deep traps caused by higher dislocation density and also due to
the increased average diameter of regions of the In uctuations
making the system less quantum-dot-like (see e.g. Ref. [343]). The
recombination via deep traps is additionally a serious concern for
non-polar GaN-based LEDs and LDs. Such non-polar devices are
considered as an alternative to polar LEDs because of the lack of the
polarization eld and respective absence of the quantum conned
Stark effect (QCSE) limiting the devices performance (see e.g. Refs.
[341,344]). However, growth of GaN in non-polar directions results
in a very high density of stacking faults and dislocations and
produces very high concentrations of deep traps that can
signicantly contribute to non-radiative recombination (see the
discussion in Ref. [185]).

In near-UV and particularly deep-UV LED structures based on


high-Al-composition ternary solutions the increased density of
deep traps, dislocations, impurities with deep levels can signicantly inuence the devices spectra and output power (see e.g. Ref.
[345]). Degradation of LED and LD performance upon operation
under high driving current is another eld where deep traps can
play an important part. We briey discuss all these issues below.
6.1. Deep traps in GaN/InGaN and AlGaN/AlGaN LEDs and LDs
It should be noted that detailed studies of deep traps spectra in
GaN-based LEDs and LDs and the attempts to correlate these
spectra with the device performance are rather scarce. To a large
degree this is related to very serious problems in acquiring and
interpreting such spectra in GaN/InGaN LED or LD structures.
Typically, the LED structure consists of some buffer layer, a severalmicrons-thick n+-GaN contact layer, the active region comprised of
a single, but more often several (usually 5) InGaN quantum wells
(QWs) with GaN barriers, a p-AlGaN electron blocking layer (EBL),
and a p-GaN contact layer. In laser structures p-AlGaN and n-AlGaN
cladding layers are added (see e.g. Ref. [346]). The ridge of the laser
diodes is several microns in width and its low capacitance is a
serious handicap for any capacitance-based measurement techniques, such as DLTS or admittance spectroscopy. Usually, to avoid
this problem testing is done on the same epitaxial structures, but
without the ridge (LED-like structures, see below).
For LED structures the attribution of certain detected traps to
this or that layer is standardly done by comparison with the results
of capacitancevoltage CV proling. But it has to be remembered
that, for the active multi-quantum-well (MQW) region, the
depletion approximation in CV proling is not strictly valid and
gives only a rough idea of the charge spatial distribution in the
QWs. One has to solve the coupled SchroedingerPoisson
equations to obtain accurate spatial distributions of the charge
on different levels in the QWs which is done by numerical
modeling and is rather time consuming (see e.g. Ref. [347]).
Moreover, the results of CV proling can be seriously affected by
the frequency of the testing signal: if the frequency is higher than
the characteristic frequency of charge exchange between the
conduction band and the QWs, the apparent CV proles can be
substantially shifted; also, at high probing frequencies, some of the
QWs can stay occupied by charge carriers even at high reverse bias
[348,349].
At low temperatures the existence of a quantum well region can
produce a series-resistance-like feature in admittance of Schottky
diodes or pn junctions with the QW region because vertical
transport through the QW requires either a temperature activation
over the barrier or tunneling (see e.g. Refs. [350,351]). In some
cases excited levels in the MQW region can produce peaks in DLTS
or admittance spectra, as well see below. The boundary of the
space charge region in LEDs can be located either in the active
MQW region or below it. In the latter case the only way to probe
defects in the MQWs is to apply forward bias or illumination.
The dominant impact of the polarization electric eld in polar
structures can sometimes result in the appearance of DLTS peaks of
the wrong sign, e.g. hole-trap-like peaks in Schottky diodes on ntype material (see below). A similar effect can be observed in LEDs.
Tunneling is often very prominent in LED structures which
provides an alternative pathway for nonequilibrium charge
relaxation in diodes. Finally, one has to remember that Mg
acceptors in the p-GaN cap of GaN/InGaN MQW LEDs are relatively
deep and strongly freeze out at temperatures below room
temperature which masks the other DLTS and admittance spectra
peaks in this temperature region and seriously complicates the
analysis of the traps level position, location and density at low
temperatures (see below). Measurements of DLTS and admittance

3.5x10

17

3.0x10

17

2.5x10

17

2.0x10

17

1.5x10

17

1.0x10

17

5.0x10

16

0.0

T=85K

1017 cm-2
QW5
QW2

virgin

QW3
1016 cm-2

0.10
Depth (m)

QW4

DLTS signal (Arb. units)

-3

Concentration (cm )

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

5*1016 cm-2
0.15

1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.4
50

45

0.6 eV

0.4 eV

1 eV
100 150 200 250 300 350 400
Temperature (K)

Fig. 82. DLTS spectrum measured on the GaN/InGaN MQW structure with reverse
bias 0.2 V, forward bias pulse of 1 V and time windows 30 ms/300 ms.

spectra on satellite GaN/InGaN MQW structures with n-GaN cap


instead of p-GaN cap and Schottky diode instead of pn junction is
sometimes helpful in understanding the essential features of deep
traps spectra.
For example, in Ref. [224] we studied electrical and luminescent
characteristics of GaN/InGaN MQW structures grown by MOCVD
on sapphire in the process similar to that used for LED structures
growth. The only differences were the substituting of the pGaN(Mg) cap in LED with undoped n-GaN cap and changing the top
ohmic contact to p-GaN in LED to the Au Schottky diode on top of nGaN cap in the test structure. CV proling of the structures
revealed the presence of 4 InGaN QWs out of 5 in the low
temperature (85 K) measurements (see Fig. 80), with the boundary
of the space charge region of the Schottky diode at 0 V located
between the rst QW1 and the second QW2 wells from the surface.
The presence of such QWs should be associated with the carriers in
the wells seeing a high barrier close to the conduction band offset
between the GaN barrier and the InGaN QW (0.7 eV in this case)
[350,351]. However, in admittance spectra of these Schottky
diodes we observed instead only very low activation energies of
20 meV and 60 meV (Fig. 81) suggesting a prominent tunneling
between the QWs. The observed proles were attributed to
decoration of the GaN/InGaN interfaces in the QWs with relatively
shallow donor traps [224]. DLTS measurements at very low reverse
bias of 0.2 V and a high forward bias pulse of 1 V showed a peak at
low temperature with the apparent activation energy of 0.4 eV and
anomalously high apparent electron capture cross section of
109 cm2, a peak with activation energy of 0.6 eV, and an unusual
hole-trap-like feature at high temperature showing the apparent
activation energy of 1 eV and the magnitude increasing with
increasing the length of the injection pulse (Fig. 82). At higher
reverse voltages the 0.4 eV and 0.6 eV signals were no longer
observed whilst the negative 1 eV peak at high temperature

became the only remaining feature (Fig. 83). The 0.4 eV peak was
attributed in Refs. [348,352] to the transition from the ground state
level in the GaN QW lled by the injection pulse to the rst excited
level in the well and subsequent tunneling into continuum. The
0.6 eV trap was attributed to the trap in the GaN barrier. The 1 eV
hole-trap-like feature in Figs. 82 and 83 was ascribed to surface
charge transition caused by switching the external voltage from
out-of-phase to in-phase with the polarization eld
[224,353]. DLTS spectra measurements with optical injection
performed for these structures indicated the presence of one
dominant hole trap peak with the activation energy of 0.9 eV
[224]. This peak was ascribed to the major acceptor level in the
GaN barriers. Such measurements were performed for structures
with the top n-GaN cap grown at 970 8C and at 1040 8C. The
abruptness of the interfaces in CV proles and the leakage current
of the structures were considerably better for the lower growth
temperature, which correlated very reasonably with the results of
varying the growth temperature of the p-GaN cap layer for
different LED structures.
In Refs. [348,352] we studied CV proles, admittance spectra,
and DLTS spectra of GaN/InGaN MQW LED structures grown by
epitaxial lateral overgrowth (ELOG) technique that allows to
radically decrease the dislocation density in the laterally overgrown regions (see e.g. Ref. [354]). As is well known the application
of this growth technique was mainly instrumental in the rst
successful demonstration of a feasible injection laser on GaN/
InGaN [355]. The structures used were true LED structures with 10mm-thick n+-GaN contact layer, 5 GaN/InGaN QWs, and 0.15 mm of
p-GaN contact layer on top, all grown by ELOG. CV proling of
these structures showed that at 0 V bias the space charge region
boundary was between the lowermost and the second lowermost
QWs. Admittance spectra of the structures (Fig. 84) showed the
presence of two major steps/peaks, one with the activation energy

450
400

2.0

350
300

G/ (pF)

80

20 meV

60 meV

60
40
20
0

100

150
200
250
Temperature (K)

300

Fig. 81. The temperature dependences of capacitance and AC conductivity G divided


by circular frequency (G/v) for the GaN/InGaN MQW structure measurements with
different frequencies from f = 70 kHz to f = 1 MHz (v = 2pf).

DLTS signal (Arb. units)

Capacitance (pF)

Fig. 80. 10 kHz CV proles measured in GaN/InGaN MQW structure at 85 K before


electron irradiation and after 10 MeV electron irradiations with uences 1016,
5  1016, and 1017 cm2.

1.5

1.1 eV

GaN/InGaN MQW
10 MeV electrons
16

1.0
0.5

5x10 cm

QW-states

-2

16

10 cm

-2

0.0
Virgin, 1 eV
-0.5
50

100 150 200 250 300 350 400


Temperature (K)

Fig. 83. DLTS spectra of the GaN/InGaN MQW structure measured before 10 MeV
electron irradiation and after irradiation with uences 1016 and 5  1016 cm2 at
reverse bias 3 V, forward bias pulse of 1 V, and time windows 500 ms/5000 ms.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Capacitance (pF)

46

600 MQW ELOG p-n CGT


500
0.185 eV (Mg)
400 0.4 eV
-related
?)
(QW
300
200
100
50

100

150 200 250


Temperature (K)

300

Fig. 84. Admittance spectrum of the ELOG MQW LED structure (the data shown for
several of the used frequencies in the 1 kHz1 MHz frequency range.

DLTS signal (Arb. units)

of 0.185 eV and the other with apparent activation energy of 0.4 eV


and very high apparent electron capture cross section. The
0.185 eV step/peak was attributed to the freeze-out of the Mg
acceptors in the p-GaN cap. After such freeze-out the apparent
position of the GaN/InGaN QW in CV proles shifted deeper inside
the sample by 0.15 mm because, at these temperatures, the p-GaN
cap behaved essentially as a dielectric layer in series with the QW
region, so that the depth calculated from the capacitance in CV
was shifted by the thickness of the p-GaN lm. DLTS spectra of the
ELOG LEDs also showed the presence of the 0.4 eV feature (Fig. 85),
alongside two prominent peaks at 0.8 eV and 1 eV. Modeling
showed that the activation energy of the 0.4 eV trap was very
close to the transition energy of 0.35 eV between the ground state
and the rst excited state in the QW from which the carrier can be
removed by tunneling. Comparison of the rest of the spectra with
DLTS spectra of ELOG n-GaN lms (Fig. 85) showed that the 0.8 and
1 eV traps are located in the n-GaN barriers. From comparison of
the two spectra in Fig. 85 one can notice the manifest absence of
the signal from the 0.2, 0.25, and 0.6 eV traps in the MQW LED DLTS
spectra, but this could easily be the consequence of interference
from the Mg freezing out stage and the 0.4 eV QW-related
transition.
These complications are not present when the deep traps
spectra in the MQW LED are studied by deep levels optical
spectroscopy (DLOS) technique [9]. (Of course, in this technique,
one has, in principle, to be concerned with the depth from which
the signal is collected. That is determined by the light absorption
coefcient. But since the measurements are mostly done in the
extrinsic excitation light range where the absorption coefcient is
low the assumption of uniform excitation within the space charge
region usually holds.) Such measurements were performed for
MQW LED structures strongly differing by the dislocation density
from 5.3  108 to 2.9  109 cm2 [356]. External quantum
efciency measurements showed a much higher value for the
lower dislocation density (see Fig. 86). The DLOS spectra of the
samples with a high and low dislocation density showed that
qualitatively the spectra were the same and were dominated by
the traps with levels near Ec1.62 eV, Ec2.11 eV, and Ec2.76 eV

10
0.2 eV

0.25 eV

0.85 eV 1 eV
0.6 eV

0.8 eV
0.1

0.4 eV

1 eV

ELOG n-GaN
ELOG MQW LED

0.01
50

100 150 200 250 300 350 400


Temperature (K)

Fig. 85. Comparison of DLTS spectra measured on ELOG n-GaN and ELOG MQW LED.

Fig. 86. External quantum efciency as a function of driving current density in


MQW LEDs with two different dislocation density.
(After Ref. [356], Fig. 1) Copyright American Institute of Physics, 2012.

Fig. 87. DLOS spectra of the two MQW LEDs with two different dislocation densities
density.
(After Ref. [356], Fig. 2) Copyright American Institute of Physics, 2012.

(Fig. 87). The traps at Ec1.62 eV and Ec2.76 eV were attributed to


the states in the InGaN QWs, the Ec2.11 eV trap was attributed to
the sates in the GaN quantum barriers (QBs) (see also Refs.
[357,358]). The absolute concentrations of traps in various parts of
the structure were determined by CV proling with illumination
with light selectively excitating only the traps in question, so called
light CV (LCV) method [9,359]. The results of this proling are
presented in Fig. 88 demonstrating the increased density of all
traps as the dislocation density increased.
DLTS spectra of MQW LEDs grown by MOCVD on Si substrates
and having very different internal quantum efciencies at low and
medium driving currents were studied by DLTS in Ref. [360]. The
authors observed two electron traps e1 and e2 with levels and
capture cross sections of respectively Ec0.33 eV, 3  1017 cm2
(e1) and Ec0.69 eV, 4  1015 cm2 (e2) in all studied samples and a
hole trap h1 (Ev + 1.04 eV, 8  1012 cm2) observed only in the
samples with high quantum efciency. The spectra for two such
samples, sample A and sample B of Ref. [360], are compared in
Fig. 89. The quantum efciency of sample A at low and medium
driving current levels was much lower than for sample B. The lower
quantum efciency correlated with the higher density of electron
traps (Fig. 90). In fact, the authors observed a linear correlation
between the concentration of electron traps e2 and the coefcient A

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

47

Fig. 88. Distribution of the tree deep traps detected by DLOS in various layers of the two MQW LEDs, solid bars refer to the high dislocation density sample, cross-hatched bars
to the low dislocation density sample.
(After Ref. [356], Fig. 4) Copyright American Institute of Physics, 2012.

in the ABC recombination model described by the equation [361]:


2

hIQE

Bn
;
An Bn2 Cn3

(11)

where hIQE is the internal quantum efciency, A is the coefcient of


non-radiative SHR recombination, B is the coefcient of radiative
recombination, C is the coefcient of Auger recombination. The e1
and e2 traps were believed to be located in the GaN close to the
active MQW region. These traps showed the logarithmic dependence of the peak magnitude in DLTS on the length of the injection
pulse and thus were attributed to point defects decorating
dislocations based on the famous work of Wosinsky [11].
Venturi et al. [362] tried to trace the deep traps in GaN/InGaN
MQW structures grown by MOCVD on sapphire to the dislocation
density in the structures by using DLTS measurements. The
dislocation density in the two structures studied in Ref. [362] were
3  108 cm2 for the low dislocation density sample LDD and
8  109 cm2 for the high dislocation density sample HDD. The
measured spectra showed two minority carriers traps A (activation
energy 0.04 eV, capture cross section 1020 to 1019 cm2) and A1
(0.120.13 eV, 1018 cm2) and a majority carrier trap B (0.5
0.54 eV, 1018 cm2) (minority carrier traps are the ones for which
the capacitance decreases after the termination of the injection
pulse, as for electron traps in p-type material, majority traps are
characterized by the increase of capacitance in DLTS transient, as
for electron traps in n-type material). The two electron traps A and
A1 showed a strong decrease of the peak magnitude on the
dislocation density in the GaN template (Fig. 91(a) and (b)) and a
logarithmic dependence of the peak magnitude on the injection

Fig. 89. DLTS spectra in MQW LEDs A and B with respectively low and high quantum
efciency.
(After Ref. [360], Fig. 2a) Copyright AIP Publishing LLC, 2014.

pulse length. These traps were, therefore, attributed to dislocation


related electron traps in p-GaN. The B trap concentration did not
vary strongly with dislocation density and the traps were ascribed
to electron traps in the GaN barrier.
As mentioned above, the possibility of using non-polar or semipolar GaN-based LED structures is actively pursued because of the
hope to suppress the detrimental effects of electronholes space
separation in polar QWs caused by the QCSE effect due to the
strong electrical polarization. However, the high density of
extended defects in non-polar structures and an increased density
of deep traps associated with these extended defects strongly
undermine the efciency of such approach. Detailed studies of
deep traps spectra and luminescence efciency in non-polar GaN
pointed to a very prominent role in this decreased efciency of
Ec0.6 eV electron traps associated with extended defects
[185]. Two other important deep traps in non-polar MQW GaN/
InGaN structures were reported in our paper [211]. The non-polar
(1010) GaN/InGaN QW structures were grown by MOCVD and
their MQW PL peak was located at 490 nm. In admittance spectra
of these structures we observed a very high density of electron
traps Ec0.41 eV located close to the QW region. These traps were
the reason of strong freeze-out of conductivity at temperatures
below room temperature. In DLTS spectra of the structures
measured above room temperature a prominent electron trap
near Ec1 eV located in the GaN barrier was observed [211]. These
two traps are believed to be responsible for the much lower
luminescence efciency of the non-polar structures compared to
their polar counterparts [211].

Fig. 90. Correlation between the trap density determined from DLTS and the A
coefcient of the non-radiative recombination in the ABC model.
(After Ref. [360], Fig. 3a) Copyright AIP Publishing LLC, 2014.

Capacitance (pF)

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

5000
4000
3000
2000
1000
0
1500

G/ (pF)

48

1000

25 meV

0.16 eV

500
75 meV
0
50 100 150 200 250 300 350 400 450
Temperature (K)

Fig. 92. Admittance spectra measured on DH AlGaN/GaN LED emitting at 360 nm.

structures with low quantum efciency and strong leakage in the


AlGaN current connement layers. The traps in question are
similar to the traps detected in GaN/InGaN LEDs and LDs after
degradation (see below).
For deep-UV structures the deep traps analysis was mostly
performed by PL and EL spectra measurements. C-related
absorption bands were found to be very detrimental for performance, as well as the VN-related PL bands in the AlGaN cladding
layers (see the sections on Mg and C doping above).
6.2. V.2.LEDs degradation studies

Fig. 91. (a) DLTS spectra in MQW LED structure with high dislocation density, (b)
the same for the sample with low dislocation density.
(After Ref. [362], Fig. 2a and b) Copyright AIP Publishing LLC, 2014.

For near-UV and deep-UV LEDs the data on deep traps is very
scarce. The near-UV LEDs are usually based on either AlGaN/GaN
MQWs or on AlGaN/GaN double heterostructures (DH). When
grown on sapphire the structures standardly use AlN/AlGaN or
AlxGa1xN/AlyGa1yN superlattices (SLs) to reduce strain and
prevent cracking. The structure is further comprised of n-AlGaN
and p-AlGaN cladding layers with the MQW region or GaN active
region in between and the p-GaN contact cap layer on top.
In Ref. [363] we presented the studies of the deep traps spectra
in near-UV AlGaN/GaN DH LEDs prepared by HVPE. HVPE growth
resulting in relatively low dislocation density in the active GaN
region allowed to switch from MQW structure to simple DH
structure. The dislocation density in high-performance LEDs
emitting near 360 nm was around 5  107108 cm2 according
to EBIC and MCL imaging [363]. The diffusion length of
nonequilibrium charge carriers determined from EBIC collection
efciency dependence on accelerating voltage [364] gave the value
of 120150 nm for the active region. This was in reasonable
agreement with such LEDs quantum efciency dependence on the
thickness of the GaN active layer as published in Ref. [365]. The
admittance spectra of the structures are shown in Fig. 92. Three
major steps/peaks corresponding to traps with activation energies
0.16 eV, 7585 MeV, and 2528 MeV were observed. The main
0.16 eV feature is related to the Mg freeze-out in the p-GaN contact
layer and in the p-AlGaN EBL. The shallow 2528 meV traps are
most likely related to oxygen donors. DLTS spectra measurements
revealed the peak due to electron traps with the level near
Ec0.35 eV, but these centers were detected only in the failed LED

Degradation of GaN/InGaN MQW blue LEDs and LDs was


studied in multiple papers. Aging tests were performed at high
driving current density and at elevated temperatures of 6070 8C
modeling the working conditions of high-power LEDs. Early work
has demonstrated for MOCVD-grown devices a clear dependence
of the degradation time on dislocation density (see e.g. Refs.
[366,367]). Decreasing the dislocation density by epitaxial lateral
overgrowth (ELOG) or by growth on low-dislocation-density GaN
substrates grown by the high-pressure technique allowed to
dramatically increase the time before degradation of lasers and
LEDs (see e.g. Ref. [355]; improved performance of low-dislocation
density LEDs was clearly illustrated in Ref. [368], for LDs see e.g.
Ref. [369]). But even devices grown by MOCVD with standard
dislocation densities demonstrate nowadays very respectable
times before degradation (thousands hours (see e.g. Ref. [370]).
Many papers point to the threshold current in lasers or
quantum efciency of LEDs to decrease with the test time t as t1/2
suggesting that the reason for degradation could be related to the
diffusion of defects into the active region of devices (see e.g. Refs.
[370372]). For MOCVD-grown LEDs and LDs with low dislocation
density grown by ELOG [369] or using low-dislocation density GaN
substrates grown by high pressure technique [371] no formation of
additional dislocations during degradation was noticed. In several
papers the effect of degradation is attributed to Mg acceptors indiffusion into the active region and forming deep trap centers (see
e.g. [369,373]). For ELOG LDs the Mg in-diffusion was directly
conrmed by secondary ion mass spectrometry (SIMS) measurements in Ref. [369].
In MBE-grown LEDs the LEDs degradation times were, until
recently, considerably shorter than in MOCVD devices. At that, the
degradation mechanisms, when studied in detail, were somewhat
different [374]. It was found that in blue MQW GaN/InGaN LEDs
with degradation times from some minutes to some 10 h two
degradation mechanisms were prevalent: (1) the formation of
dislocation bunches in the (0 0 0 1) plane (and associated increase
in non-radiative recombination), and (2) increased nonuniformity
of injection in pn junctions [374].

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

49

Fig. 93. (a) Evolution of optical power at different currents for GaN/InGaN MQW LDs after different stress times, (b) the evolution of the lasers threshold current with stress
time (the black line is the t1/2 approximation, the red line is the erfc approximation), (c) optical power of LED structures on the same wafers as the LDs measured as a function
of driving current for several stress times; (d) the LEDs optical power degradation with stress time for different driving currents. (For interpretation of the references to color
in this gure legend, the reader is referred to the web version of this article.)
(After Ref. [370], Fig. 1ad) Copyright American Institute of Physics, 2010.

For MOCVD LDs and LEDs perhaps the most insightful studies
were reported in papers [370,372]. The authors performed
measurements of degradation parameters in MQW LDs and
LED-like structures, both prepared on the same epitaxial wafers
and differing by the presence of the ridge and mirror facets in lasers
(absent in LED structures). It was found that the degradation of
electrical and optical parameters during the stress test at the same
forward driving current density of 4 kA/cm2 at 75 8C occurred in a
similar fashion for LEDs and LDs, thus allowing to exclude the laser
facets degradation as the main reason for the lasers failure after
stress test. Rather, the reason had to be related to the increased
impact of non-radiative recombination in the active region of both
types of structures. The authors observed the increase of the
threshold current of LDs and the decrease of the quantum
efciency of LEDs with the test time occurring as the square root
of time or, more precisely, as a complementary error function (erfc)
expected to describe the diffusion with the diffusion coefcient D.
It was observed that, as it should be, the degradation of the optical
output for LEDs was more pronounced for the lower driving

Fig. 94. Evolution of LEDs IV characteristics with stress time.


(After Ref. [370], Fig. 3) Copyright American Institute of Physics, 2010.

currents region where the SRH deep traps impact on non-radiative


recombination is stronger. These conclusions are illustrated by
Fig. 93. The diffusion coefcient for the defect diffusion estimated
from these measurements was quite high, D = (15)  1019 cm2/s.
It was also observed that the degradation of optical power was
closely related to the increase of the defect-related low-voltage
forward current leakage in LEDs attributed to the loss of injection
efciency caused by the increased component of non-injection
tunneling current (Fig. 94). A similar effect was reported for
degradation of blue LEDs in Refs. [375378] where it was
attributed to enhanced tunneling via defect levels on dislocations.
The defect species responsible for the LEDs/LDs degradation
described in Ref. [370] and illustrated by Figs. 93 and 94 was more
denitively pinpointed in DLTS measurements performed on LEDs
before and after degradation (see Ref. [372]; DLTS measurements
on LDs were not possible because of the low capacitance of the
structure, but the dominant defects should be the same in both
structures given the same degradation characteristics, as demonstrated by Fig. 93). The said DLTS spectra of the studied structures
are presented in Fig. 95 for different stress test times. One can see a
very broad feature belonging to electron traps located in the MQW
region and having the apparent activation energy of 0.350.45 eV.
The concentration of these traps increased with time as t1/2 and the
changes in the traps concentration obviously correlated with the
increase of the threshold current in lasers and the decrease of the
optical power output in LEDs [372]. Thus, these traps seem to be
responsible for the observed degradation. The possible nature of
the traps is not totally clear. The authors of Ref. [372] argue that
this trap should be similar to the Ec0.6 eV trap in n-GaN if one
considers that the trap is located in the InGaN QW and its energy
should be lower than in GaN by approximately DEc, the conduction
bands offset (the structures in Refs. [370,373] emitted at 405 nm).
The origin of different traps at Ec0.6 eV in n-GaN has been
discussed above in the section on deep traps in GaN. It seems likely
that these traps are related to extended defects decorated by native
defects, although other interpretations have been proposed.

50

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Fig. 95. DLTS spectra in LEDs measured after different stress times.
(After Ref. [372], Fig. 2b) Copyright American Institute of Physics, 2011.

Notably, the center with the level near Ec0.6 eV has been
attributed in several papers to Mg complexes with nitrogen
vacancies (see e.g. Refs. [216,379]) which would tie neatly the
degradation behavior with the reported observations of Mg indiffusion. But more study is obviously necessary here. It should be
perhaps noted that the traps behavior upon degradation seems to
be similar to the behavior of QW-related traps upon 10 MeV
irradiation reported by us in Ref. [224]. The studied structures were
the ones for which the starting deep traps spectra and CV
concentration proles are presented in Figs. 8083. Electron
irradiation created traps with energies 0.1 and 0.2 eV decorating
the GaN/InGaN QW interfaces. As the result effective compensation of the MQW region occurred as shown in Fig. 80 for several
10 MeV electrons uences. Simultaneously, in DLTS spectra
appeared and grew in amplitude a broad feature related to
trapping in the MQW region (Fig. 83). This feature is clearly
reminiscent of the 0.350.45 eV trap feature in degraded LEDs in
Fig. 95. In tune with that we observed the emergence of a
prominent electron traps peak with activation energy 1.1 eV
whose concentration increased with electron uence. These traps
most likely were formed in n-GaN barriers of the MQW region. The
logarithmic dependence of their peak magnitude on the injection
pulse suggests that these traps belong to dislocations decorated
with point defects, very likely nitrogen interstitials [224]. The
electron traps of the said type also dominated the deep traps
spectra measured by DLTS with optical excitation. At electron
uences higher than 1016 cm2 they totally suppressed the signal
from the hole traps near Ev + 0.9 eV in the n-GaN barriers. The deep
electron traps formation after electron bombardment with
uences higher than 1016 cm2 led to very substantial decrease
of the MQW-related luminescence efciency.
Degradation of blue GaN/InGaN MQW LEDs prepared by
MOCVD and near-UV AlGaN/GaN DH LEDs grown by HVPE was
studied by Shmidt et al. [375]. The authors observed a somewhat
similar behavior in both cases: an increase with aging of the
forward current and reverse current attributed to enhanced
tunneling via dislocations decorated by point defects. This creates
a tunneling shunt not contributing to injection and thus decreasing
the internal quantum efciency of devices. The formation of such
shunts was ascribed to local heating of the structures causing In or
Al segregation on dislocations [378]. The lifetime of the near-UV
AlGaN/GaN DH LEDs was found to be considerably lower than for

Fig. 96. (a) Electroluminescence spectra of 285-nm LEDs as function of aging time.
(After Ref. [385], Fig. 6) Copyright American Institute of Physics, 2011.

blue GaN/InGaN MQW LEDs, even though the dislocation density


in the AlGaN/GaN structures was not higher [375].
For deep-UV MQW AlGaN/AlGaN LEDs, effects of degradation
caused by prolonged operation under high forward current of
100 mA were studied in several papers. Catastrophic degradation
observed was attributed to the formation of macroscopic defects
[380] or to severe current crowding [381]. Gradual degradation
reported in Refs. [382384] was not related to increased nonradiative recombination in the MQW region, but rather ascribed to
generation of defects in the AlGaN cladding layers. The effect was
variously explained by migration of Al in the p-AlGaN cladding
layer or to enhanced compensation of p-type conductivity in this
layer. Perhaps, the most detailed studies so far have been described
in Ref. [385] for the MQW LED structures emitting at 310 nm or
285 nm. Time resolved photoluminescence spectra kinetics
measurements showed that no degradation of the PL or EL
quantum efciency occurred in the MQW active region. The overall
EL efciency was degrading more rapidly than for blue LEDs
(typical degradation times of some 10 h) and was accompanied by
the increased intensity of the long wavelength recombination band
in the spectra (Fig. 96). Comparison of the EL spectra before and
after degradation with the PL spectra selectively excited in the
MQW region showed that the long wavelength band originated in
the p-AlGaN cladding layer (Fig. 97). This band can be credibly
attributed to the recombination involving VN donors (see above the
sections devoted to Mg and C behavior in AlGaN). These
degradation effects occurred simultaneously with the strong

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

51

Table 5
Electron and hole traps detected in MOCVD-grown GaN/InGaN MQWs LEDs and LDs.
Trap name and type
e1, electron
e2, electron
h1, hole
A, electron
A1, electron
B, electron
Electron
Electron
Electron
Electron
Hole
Electron
Electron
a
b
c
d
e

Trap level (eV)


Ec0.33
Ec0.69
Ev+1.04
Ec0.04
Ec0.12
Ec0.54
Ec(0.350.45)
Ec0.19
Ec0.6
Ec1.1
Ev+0.9
Ec0.41
Ec1

Capture cross section (cm2)


17

3  10
4  1015
8  1012
1019
1018
1018

2.6  1014
1012
1012
1012

Location in the structure


Near QW
GaN barrier
p-GaN?
p-GaN
GaN barrier
QW
Interface
GaN barrier
GaN barrier
GaN barrier
QW?
GaN barrier

Dislocation related?
a

Yes
Yesa
?
Yesb
Yesb
?
?
?
?
Yesa
?
Yes
Yesa

Reference
DLTS [360]
DLTS [360] c
DLTS [360]
DLTS [362]
DLTS [362]
DLTS [362]
DLTS [372] c
AS [224] d
DLTS [352]
DLTS, ODLTS [224] d
ODLTS [224]
AS [211] e
DLTS [211] e

Determined from the logarithmic DLTS peak amplitude dependence on injection pulse length.
Correlation with dislocation density and logarithmic dependence of the DLTS peak.
Degrades optical output.
Produced by electrical stress, determines output degradation electron irradiation.
Observed in non-polar LEDs with high density of extended defects.

increase in the low-voltage forward current region of IV


characteristics explained by the formation of the tunneling shunt
due to defect-decorated dislocations. This tunneling current
component decreased the injection efciency at the given forward
current and thus the density of non-equilibrium charge carriers in
the active MQW region. This decrease in the injection efciency

was additionally contributed to by the increased compensation of


p-type conductivity in p-AlGaN cladding. The latter effect is
believed to be caused by the introduction of compensating
nitrogen vacancy donors.
Finally, a few words about radiation-induced changes in GaNLEDs. The data on such changes is scarce and not systematic. We do
not consider in detail the LEDs degradation studies after irradiation
in this survey. The results existing in the literature have been
reviewed recently in Refs. [17,18,323]. These data were produced
mostly in the early days of GaN devices studies and are very
fragmentary. The general conclusion is that the radiation tolerance
of GaN-based LEDs is at least an order of magnitude higher than for
their GaAs counterparts and there is no immediate need to be
concerned about practical aspects of performance degradation of
GaN-based LEDs in space applications. However, radiation
degradation experiments can provide useful clues to the reasons
of LEDs and LDs degradation under electrical stress. From this
perspective, it seems such experiments should be carried out on a
much larger scale than up to now.
In Table 5 we summarized what is known about the deep traps
in GaN-based LED and LD structures. It seems that, in GaN/InGaN
MQW LEDs, deep traps near Ec0.04, Ec0.14, Ec0.4 eV, Ec0.6,
Ec1.1 eV could be traced to higher dislocation density. Among
them the Ec0.4 eV, Ec0.6 eV, and Ec1.1 eV seem to play
important role in decreasing the LEDs optical output. They also
seem to be at least partly responsible for devices performance
degradation after aging tests and after electron irradiation.
7. Conclusions

Fig. 97. (a) EL and PL spectra in 285 nm LEDs before aging, excitation of only QWs in
PL measurements, (b) the same as (a) after aging, (c) EL and PL spectra with PL
excitation of the top p-AlGaN cladding layer (After Ref. [385], Fig. 7ac) Copyright
American Institute of Physics, 2011.

To conclude this discussion it needs to be said that recent


theoretical calculations seem to strongly support the idea that
carbon in GaN and AlGaN is a deep acceptor causing strong
compensation and high resistivity when present in sufcient
concentration. This deep C acceptor plays, at least in some cases,
important role in determining the intensity of yellow-luminescence-like recombination in GaN and strong defect absorption in
AlN. The behavior of Mg acceptors in GaN and AlGaN seems to be
strongly determined by compensation by native donors, presumably, VN-related. Fe doping of GaN and AlGaN provokes the
introduction of deep traps deep in the bandgap. Some of the
experimental data suggests that the relevant level in GaN could be
close to Ec0.5 eV and to behave as a center with the level tied to
the level of vacuum in AlGaN. At the same time Fe doping often
induces the formation of not-Fe-related deep acceptor near
Ec0.6 eV very likely associated with extended defects in GaN.

52

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

Such centers play important role in gate lag and drain lag observed
in AlGaN/GaN HEMTs where they seem to be located near the
AlGaN/GaN interface on the GaN side. The majority of deep
electron traps in GaN are most likely dislocation-related and that
probably explains the prominent role of dislocations in the
trapping, gate leakage, subthreshold current leakage, and degradation in AlGaN/GaN HEMTs. The gate leakage in AlGaN/GaN
HEMTs seems to be promoted by open-core screw dislocations and,
in InAlN/GaN HEMTs, by dislocations decorated by In. Some major
hole traps in n-GaN and n-AlGaN are most likely related to gallium
vacancies complexes with Si or oxygen and to C, which reasonably
explains the behavior of YL-like, GL-like, and RL-like PL/MCL bands
in AlGaN. The blue luminescence band attribution seems to require
either the presence of electron traps in the upper half of the
bandgap or the (VGa2O) hole traps in the lower half of the
bandgap. Each of these centers, depending on relative concentrations, can play outstanding role in hole trapping determining the
charge collection efciency and photosensitivity of GaN-based
radiation detectors and photodetectors. The light emission
efciency of GaN/InGaN MQW LEDs and the threshold current
in GaN/InGaN MQW LDs seem to be greatly affected by deep
electron traps near Ec(0.30.6) eV located close to or in the MQW
active region. The traps are most likely dislocation-related and
such dislocation-related centers play important role in LEDs and
LDs degradation after electrical stress. In many cases one observes
similar effects upon stress tests degradation of HEMTs and LEDs
with the changes induced by high energy particles irradiation. The
work along these directions is at the very beginning and much
more study is necessary to clarify the nature of the observed
effects, but they are denitely very important for optimization of
GaN-based devices performance.
Acknowledgments
The authors would like to gratefully acknowledge the contribution to this paper of long standing collaboration with Dr. N.B.
Smirnov at MISiS (Moscow), Prof. E. Yakimov (IPTM, Chernogolovka,
Russia), Prof. S.J. Pearton, Prof. Fan Ren (both University of
Florida, Gainesville USA), of Prof. H. Amano at Nagoya University
(Nagoya, Japan), Prof. J. Han at Yale University (New Haven, USA),
and Dr. A. Dabiran at SVT Technologies (Minneapolis, USA). The work
at NUST MISiS was supported in part by the Ministry of Education
and Science of the Russian Federation in the framework of Increased
Competitiveness Program of NUST MISiS (No. K2-2014-055). The
work at Chonbuk National University was supported by National
Research Foundation of Korea (NRF) funded by Ministry of Science,
ICT & Future Planning (2013R1A2A2A07067688, 2010-0019626).
AYP would like to acknowledge support from the Brain Pool program
of the Korean Government during his stay at Chonbuk National
University.
References
[1] D.V. Lang, J. Appl. Phys. 45 (1974) 30233032.
[2] G.M. Martin, A. Mitonneau, D. Pons, A. Mircea, D.W. Woodard, J. Phys. C: Solid
State Phys. 13 (1980) 38553882.
[3] C. Hurtes, M. Boulou, A. Mitonneau, D. Bois, Appl. Phys. Lett. 32 (1978) 821823.
[4] J.K. Rhee, P.K. Bhattacharya, J. Appl. Phys. 53 (1982) 42474249.
[5] R.E. Kremer, M.C. Arikan, J.C. Abele, J.S. Blakemore, J. Appl. Phys. 62 (1987) 2424
2431.
[6] J.C. Abele, R.E. Kremer, J.S. Blakemore, J. Appl. Phys. 62 (1987) 24322438.
[7] Peter K. Schroeder, Semiconductor Material and Device Characterization, WileyInterscience Int., New York, 1990.
[8] J.L. Pautrat, B. Katircioglu, N. Magnea, D. Bensahel, J.C. Pster, L. Revoil, SolidState Electron. 23 (1980) 1159.
[9] P. Blood, J.W. Orton, The Electrical Characterization of Semiconductors: Majority
Carriers and Electron States, Academic Press, San Diego, 1992.
[10] A. Chantre, G. Vincent, D. Bois, Phys. Rev. B 23 (1981) 53355339.
[11] T. Wosinsky, J. Appl. Phys. 65 (1989) 1566.
[12] C.G. Van de Walle, J. Neugebauer, J. Appl. Phys. 95 (2004) 3851.

[13] D.C. Look, Z.-Q. Fang, B. Clain, J. Cryst. Growth 281 (2005) 143.
[14] M.A. Reshchikov, H. Morkoc, J. Appl. Phys. 97 (2005) 061301.
[15] D.W. Palmer, in: P. Bhattacharya, R. Fornari, H. Kamimura (Eds.), Comprehensive Semiconductor Science and Technology, vol. 4, Elsevier, Amsterdam, 2011,
pp. 390447.
[16] D.C. Look, Phys. Stat. Sol. B 228 (2001) 293.
[17] A.Y. Polyakov, in: S.J. Pearton (Ed.), GaN and ZnO-based Materials and Devices,
Springer Series in Materials Science, Springer, Heidelberg, 2012, pp. 251294.
[18] S.J. Pearton, R. Deist, F. Ren, L. Liu, A. Polyakov, J. Kim, J. Vac. Sci. Technol. A 31
(2013) 050801.
[19] D.W. Jenkins, J.D. Dow, M.H. Tsai, J. Appl. Phys. 72 (1992) 4130.
[20] J. Neugebauer, C.G. Van de Walle, Phys. Rev. B 50 (1994) 80678070.
[21] J. Neugebauer, C.G. Van de Walle, Appl. Phys. Lett. 69 (1996) 503505.
[22] T. Mattila, R.M. Nieminen, Phys. Rev. B 55 (1997) 95719576.
[23] P. Boguslawski, E.L. Briggs, J. Bernholc, Appl. Phys. Lett. 69 (1996) 233.
[24] I. Gorczyca, A. Svane, N.E. Christensen, Phys. Rev. B 60 (1999) 8147.
[25] C. Stamp, C.G. Van de Walle, D. Vogel, P. Kruger, J. Pollmann, Phys. Rev. B 61
(2000) R7846R7849.
[26] C. Stamp, C.G. Van de Walle, Phys. Rev. B 65 (2002) 155212.
[27] M.A. Reshchikov, H. Morkoc, J. Appl. Phys. 97 (2005) 061301.
[28] C.H. Park, D.J. Chadi, Phys. Rev. B 55 (1997) 3028.
[29] C.G. Van de Walle, Phys. Rev. B 57 (1998) R2033R2036.
[30] D.J. Chadi, K.J. Chang, Phys. Rev. Lett. 61 (1988) 873.
[31] P. Bogusawski, J. Bernholc, Phys. Rev. B 56 (1997) 94969505.
[32] S.J. Pearton, A.Y. Polyakov, Chem. Vap. Depos. 16 (2010) 266274.
[33] J. Neugebauer, C.G. Van de Walle, Phys. Rev. Lett. 75 (1995) 4452.
[34] S. Limpijumnong, C.G. Van deWalle, Phys. Stat. Sol. B 228 (2001) 303307.
[35] J. Neugebauer, C.G. Van de Walle, Appl. Phys. Lett. 68 (1996) 1829.
[36] A.F. Wright, Phys. Rev. B 60 (1999) R5101R5104.
[37] S.M. Myers, A.F. Wright, G.A. Petersen, W.R. Wampler, C.H. Seager, M.H. Crawford, J. Han, J. Appl. Phys. 89 (2001) 31953202.
[38] S.M. Myers, A.F. Wright, G.A. Petersen, C.H. Seager, W.R. Wampler, M.H. Crawford, J. Han, J. Appl. Phys. 88 (2000) 46764687.
[39] J.L. Lyons, A. Janotti, C.G. Van de Walle, Phys. Rev. Lett. 108 (2012) 156403.
[40] C. Van de Walle, J. Neugebauer, Nature 423 (2003) 626628.
[41] A.F. Wright, J. Appl. Phys. 92 (2002) 25752585.
[42] J.L. Lyons, A. Janotti, C.G. Van de Walle, Appl. Phys. Lett. 97 (2010) 152108.
[43] J.L. Lyons, A. Janotti, C.G. Van de Walle, Phys. Rev. B 89 (2014) 035204.
[44] M. Himmerlich, A. Knubel, R. Aidam, L. Kirste, A. Eisenhardt, S. Krischok, J.
Pezoldt, P. Schley, E. Sakalauskas, R. Goldhahn, R. Felix, J.M. Manuel, F.M.
Morales, D. Carvalho, T. Ben, R. Garca, G. Koblmuller, J. Appl. Phys. 113
(2013) 033501.
[45] Q. Yan, A. Janotti, M. Schefer, C.G. Van de Walle, Appl. Phys. Lett. 100 (2012)
142110.
[46] D.O. Demchenko, I.C. Diallo, M.A. Reshchikov, Phys. Rev. Lett. 110 (2013) 087404.
[47] J.W. Ager III, R.E. Jones, D.M. Yamaguchi, K.M. Yu, W. Walukiewicz, S.X. Li, E.E.
Haller, H. Lu, W.J. Schaff, Phys. Stat. Sol. B 244 (2007) 18201824.
[48] J. Elsner, R. Jones, P.K. Sitch, V.D. Porezag, M. Elstner, Th. Frauenheim, M.I. Heggie,
S. Oberg, P.R. Briddon, Phys. Rev. Lett. 79 (1997) 3672.
[49] R. Jones, J. Elsner, M. Haugk, R. Gutierrez, T. Frauenheim, M.I. Heggie, S. Uberg,
P.R. Briddon, Phys. Stat. Sol. A 171 (1999) 167173.
berg, P.R.
[50] J. Elsner, R. Jones, M.I. Heggie, P.K. Sitch, M. Haugk, Th. Frauenheim, S. O
Briddon, Phys. Rev. B 58 (1998) 12571.
[51] D.C. Look, J.R. Sizelove, Phys. Rev. Lett. 82 (1999) 12371240.
[52] J.H. You, J.-Q. Lu, H.T. Johnson, J. Appl. Phys. 99 (2006) 033706.
[53] J.E. Northrup, Appl. Phys. Lett. 78 (2001) 22882290.
[54] M. Lee, M.A. Belkhir, X.Y. Zhu, Y.H. Lee, Y.G. Hwang, T. Frauenheim, Phys. Rev. B
61 (2000) 1603316039.
[55] A.F. Wright, U. Grossner, Appl. Phys. Lett. 73 (1998) 2751.
[56] K.C. Mishra, K.H. Johnson, P.C. Schmidt, Phys. Stat. Sol. A 208 (2011) 1555
1557.
[57] L. Lymperakis, M. Albrecht, T. Remmele, H.P. Strunk, J. Neugebauer, Phys. Rev.
Lett. 93 (2004) 196401.
[58] H. Lei, J. Chen, P. Ruterana, Appl. Phys. Lett. 96 (2010) 161901.
[59] F. Tuomisto, S. Hautakangas, I. Makkonen, V. Ranki, M.J. Puska, K. Saarinen, M.
Bockowski, T. Suski, T. Paskova, B. Monemar, X. Xu, D.C. Look, Phys. Stat. Sol. B
243 (2006) 14361440.
[60] L. Chang, S.K. Lai, F.R. Chen, J.J. Kai, Appl. Phys. Lett. 79 (2001) 928.
[61] A. Minj, D. Cavalcoli, A. Cavallini, Appl. Phys. Lett. 97 (2010) 132114.
[62] J. Song, F.J. Xu, X.D. Yan, F. Lin, C.C. Huang, L.P. You, T.J. Yu, X.Q. Wang, B. Shen, K.
Wei, X.Y. Liu, Appl. Phys. Lett. 97 (2010) 232106.
[63] B.S. Simpkins, E.T. Yu, P. Waltereit, J.S. Speck, J. Appl. Phys. 94 (2003) 1448.
[64] A. Lochthofen, W. Mertin, G. Bacher, L. Hoeppel, S. Bader, J. Off, B. Hahn, Appl.
Phys. Lett. 93 (2008) 022107.
[65] E. Valcheva, T. Paskova, B. Monemar, J. Cryst. Growth 255 (2003) 1926.
[66] T.B. Wei, R.F. Duan, J.X. Wang, J.M. Li, Z.Q. Huo, Y.P. Zeng, Microelectron. J. 39
(2008) 15561559.
[67] A.Y. Polyakov, E.B. Yakimov, N.B. Smirnov, A.V. Govorkov, A.S. Usikov, H. Helava,
Y.N. Makarov, I.-H. Lee, J. Vac. Sci. Technol. B 32 (2014) 051212.
[68] B. Kim, D. Moon, K. Joo, S. Oh, Y.K. Lee, Y. Park, Y. Nanishi, E. Yoon, Appl. Phys. Lett.
104 (2014) 102101.
[69] S. Bychikhin, D. Poganya, L. Vandamme, G. Meneghesso, E. Zanoni, J. Appl. Phys.
97 (2005) 123714.
[70] W. Gotz, R.S. Kern, C.H. Chen, H. Liu, D.A. Steigerwald, R.M. Fletcher, Mat. Sci. Eng.
B 59 (1999) 211.
[71] P.P. Paskov, et al. Phys. Stat. Sol. C 6 (2009) S763.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156
[72] C. Wetzel, T. Suski, J.W. Ager, E.R. Weber, E.E. Haller, S. Fisher, B.K. Meyer, R.J.
Molnar, P. Perlin, Phys. Rev. Lett. 78 (1997) 3923.
[73] M.D. McCluskey, N.M. Johnson, C.G. Van de Walle, D.P. Bour, M. Kneissl, W.
Walukiewicz, Phys. Rev. Lett. 80 (1998) 40084011.
[74] A.Y. Polyakov, M. Shin, J.A. Freitas, M. Skowronski, D.W. Greve, R.G. Wilson, J.
Appl. Phys. 80 (1996) 63496354.
[75] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, M.G. Milvidskii, J.M. Redwing, M.
Shin, M. Skowronski, D.W. Greve, R.G. Wilson, Solid-State Electron. 42 (1998)
627.
[76] Y. Taniyasu, M. Kasu, N. Kobayashi, Appl. Phys. Lett. 81 (2002) 1255.
[77] M.L. Nakarmi, K.H. Kim, K. Zhu, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 85 (2004)
3769.
[78] R. Zeisel, M.W. Bayerl, S.T.B. Goennenwein, R. Dimitrov, O. Ambacher, M.S.
Brandt, M. Stutzmann, Phys. Rev. B 61 (2000) R16283R16286.
[79] T. Ive, O. Brandt, H. Kostial, K.J. Friedland, L. Daweritz, K.H. Ploog, Appl. Phys. Lett.
86 (2005) 024106.
[80] N.T. Son, M. Bickermann, E. Janzen, Appl. Phys. Lett. 98 (2011) 092104.
[81] X.T. Trinh, D. Nilsson, I.G. Ivanov, E. Janzen, A. Kakanakova-Georgieva, N.T. Son,
Appl. Phys. Lett. 103 (2013) 042101.
[82] R. Butte, J.-F. Carlin, E. Feltin, M. Gonschorek, S. Nicolay, G. Christmann, D.
Simeonov, A. Castiglia, J. Dorsaz, H.J. Buehlmann, S. Christopoulos, G. Baldassarri
Hoger von Hogersthal, A.J.D. Grundy, M. Mosca, C. Pinquier, M.A. Py, F. Demangeot, J. Frandon, P.G. Lagoudakis, J.J. Baumberg, N. Grandjean, J. Phys. D: Appl.
Phys. 40 (2007) 6328.
[83] M.A. Py, L. Lugani, Y. Taniyasu, J.-F. Carlin, N. Grandjean, Phys. Rev. B 90 (2014)
115208.
[84] A.Y. Polyakov, I.-H. Lee, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, S.J.
Pearton, J. Appl. Phys. 109 (2011) 123701.
[85] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, V.M.
Zaletin, I.M. Gazizov, N.G. Kolin, S.J. Pearton, J. Vac. Sci. Technol. B 30 (2012)
021205.
[86] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, A.S. Usikov, H. Helava, Y.N. Makarov, S.J.
Pearton, J. Appl. Phys. 115 (2014) 223702.
[87] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, N.G.
Kolin, V.M. Boiko, A.V. Korulin, S.J. Pearton, J. Vac. Sci. Technol. B 29 (2011)
041201.
[88] K.B. Nam, M.L. Nakarmi, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 86 (2005) 222108.
[89] N. Nepal, M.L. Nakarmi, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 89 (2006) 092107.
[90] A. Sedhain, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 100 (2012) 221107.
[91] B. Bastek, F. Bertram, J. Christen, T. Hempel, A. Dadgar, A. Krost, Appl. Phys. Lett.
95 (2009) 032106.
[92] H. Nakayama, P. Hacke, M.R.H. Khan, Jpn. J. Appl. Phys. 2 35 (1996) L282.
[93] W. Gotz, N.M. Johnson, J. Walker, D.P. Bour, A. Street, Appl. Phys. Lett. 68 (1996)
667.
[94] T.D. Moustakas, R.J. Molnar, Mater. Res. Soc. Symp. Proc. 281 (1993) 753.
[95] W. Kim, A. Salvador, A.E. Botchkarev, O. Atkas, S.N. Mohammad, H. Morcoc, Appl.
Phys. Lett. 69 (1996) 559.
[96] T. Tanaka, A. Watanabe, H. Amano, Y. Kobayashi, I. Akasaki, S. Yamazaki, M.
Koike, Appl. Phys. Lett. 65 (1994) 593.
[97] T. Mori, K. Nagamatsu, K. Nonaka, K. Takeda, M. Iwaya, S. Kamiyama, H. Amano, I.
Akasaki, Phys. Stat. Sol. C 6 (2009) 26212625.
[98] S.-R. Jeon, Z. Ren, G. Cui, J. Su, M. Gherasimova, J. Han, H.-K. Cho, L. Zhou, Appl.
Phys. Lett. 86 (2005) 082107.
[99] P.R. Tavernier, T. Margalith, J. Williams, D.S. Green, S. Keller, S.P. Denbaars, U.K.
Mishra, S. Nakamura, D.R. Clarke, J. Cryst. Growth 264 (2004) 150.
[100] Z. Lillental Weber, M. Benamara, W. Swider, J. Washburn, I. Grzegory, S. Porowski, R.D. Dupuis, C.J. Eiting, Mater. Res. Soc. Symp. Proc. 595 (1999) W9.7.1.
[101] A. Bell, R. Liu, F.A. Ponce, H. Amano, I. Akasaki, D. Cherns, Appl. Phys. Lett. 349
(2003) 82.
[102] A.K. Rice, K.J. Malloy, J. Appl. Phys. 89 (2001) 28162825.
[103] N.M. Shmidt, A.G. Kolmakov, A.V. Loskutov, A.Y. Polyakov, N.B. Smirnov, A.V.
Govorkov, S.J. Pearton, A.V. Osinsky, Solid-State Electron. 47 (2003) 10031008.
[104] I.D. Goepfert, E.F. Schubert, A. Ossinski, P.E. Norris, N.N. Faleev, J. Appl. Phys. 18
(2000) 2030.
[105] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Osinsky, P.E. Norris, S.J. Pearton,
J. Van Hove, A. Wowchak, P. Chow, J. Appl. Phys. 90 (2001) 40324038.
[106] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.M. Dabiran, A.V. Osinsky, S.J.
Pearton, Appl. Phys. Lett. 89 (2006) 2354443.
[107] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, K.D. Shcherbatchev, V.T. Bublik, M.I.
Voronova, A.M. Dabiran, A.V. Osinsky, S.J. Pearton, J. Vac. Sci. Technol. B 25
(2007) 6973.
[108] J.W. Huang, T.F. Kuech, H. Lu, I. Bhat, Appl. Phys. Lett. 68 (1996) 2392.
[109] D.J. Kim, D.Y. Ryu, N.A. Bojarczuk, J. Karasinski, S. Guha, S.H. Lee, J.H. Lee, J. Appl.
Phys. 88 (2000) 25642569.
[110] D.J. Kim, J. Appl. Phys. 88 (2000) 19291934.
[111] M. Schmeits, N.D. Nguyen, M. Germain, J. Appl. Phys. 89 (2001) 18901897.
[112] D. Seghier, H.P. Gislason, J. Appl. Phys. 88 (2000) 64836487.
[113] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, K.H. Baik, S.J. Pearton, B. Luo, F. Ren,
J.M. Zavada, J. Appl. Phys. 94 (2003) 39603965.
[114] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, S.J. Pearton, J.M. Zavada, R.G. Wilson,
J. Appl. Phys. 94 (2003) 30693074.
[115] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, A.M. Dabiran, P.P.
Chow, A.M. Wowchak, I.-H. Lee, J.-W. Ju, S.J. Pearton, J. Appl. Phys. 106 (2009)
073706.
[116] A.Y. Polyakov, A.V. Govorkov, N.B. Smirnov, A.E. Nikolaev, I.P. Nikitina, V.A.
Dmitriev, Solid-State Electron. 45 (2001) 261265.

53

[117] A. Usikov, O. Kovalenkov, V. Soukhoveev, V. Ivantsov, A. Syrkin, V. Dmitriev, A.Y.


Nikiforov, S.G. Sundaresan, S.J. Jeliazkov, A.V. Davydov, Phys. Stat. Sol. C 5 (2008)
18291831.
[118] V. Dmitriev, A. Usikov, in: Z.C. Feng (Ed.), III-Nitride Semiconductor Materials,
Imperial College Press, London, 2006, pp. 140.
[119] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, N.G. Kolin, D.I. Merkurisov, V.M. Boiko, K.D. Shcherbatchev, V.T. Bublik, M.I. Voronova, S.J. Pearton,
A. Dabiran, A.V. Osinsky, J. Vac. Sci. Technol. B 24 (2006) 22562261.
[120] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, S.J. Pearton, J.M. Zavada, J. Phys.:
Condens. Matter 16 (2004) 16.
[121] B.I. Shklovski, A.L. Efros, Nauka (Moscow) (1979) (in Russian).
[122] B. Monemar, P.P. Paskov, G. Pozina, et al. Phys. Rev. Lett. 102 (2009) 235501.
[123] B. Monemar, P.P. Paskov, J.P. Bergman, et al. Phys. Stat. Sol. B 248 (2006) 1604.
[124] N. Nepal, K.B. Nam, M.L. Nakarmi, J.Y. Lin, H.X. Jiang, J.M. Zavada, R.G. Wilson,
Appl. Phys. Lett. 84 (2004) 1090.
[125] M.L. Nakarmi, N. Nepal, C. Ugolini, T.M. Altahtamouni, J.Y. Lin, H.X. Jiang, Appl.
Phys. Lett. 89 (2006) 152120.
[126] M.L. Nakarmi, N. Nepal, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 94 (2009) 091903.
[127] B.E. Gaddy, Z. Bryan, I. Bryan, R. Kirste, J. Xie, R. Dalmau, B. Moody, Y. Kumagai, T.
Nagashima, Y. Kubota, T. Kinoshita, A. Koukitu, Z. Sitar, R. Collazo, D.L. Irving,
Appl. Phys. Lett. 103 (2013) 161901.
[128] T. Kinoshita, T. Obata, H. Yanagi, S.-I. Inoue, Appl. Phys. Lett. 102 (2013) 012105.
[129] Y. Taniyasu, M. Kasuano, T. Makimoto, Nature 441 (2006) 325.
[130] J. Li, Z.Y. Fan, R. Dahal, M.L. Nakarmi, J.Y. Lin, H.X. Jiang, Appl. Phys. Lett. 89 (2006)
213510.
[131] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, K.H. Baik, S.J. Pearton, B. Luo, F. Ren,
J.M. Zavada, J. Vac. Sci. Technol. B 22 (2004) 771775.
[132] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, K.H. Baik, S.J. Pearton, J.M. Zavada, J.
Vac. Sci. Technol. B 22 (2004) 22912294.
[133] C.H. Seager, A.F. Wright, J. Yu, W. Gotz, J. Appl. Phys. 92 (2002) 6553.
[134] A. Armstrong, A.R. Arehart, B. Moran, S.P. DenBaars, U.K. Mishra, J.S. Speck, S.A.
Ringel, Appl. Phys. Lett. 84 (2004) 374.
[135] C.H. Seager, D.R. Tallant, J. Yu, W. Gotz, J. Lumin. 106 (2004) 115124.
[136] D.S. Green, U.K. Mishra, J.S. Speck, J. Appl. Phys. 95 (2004) 8456.
[137] A. Armstrong, C. Poblenz, D.S. Green, U.K. Mishra, J.S. Speck, S.A. Ringel, Appl.
Phys. Lett. 88 (2006) 082114.
[138] A. Armstrong, A.R. Arehart, D. Green, U.K. Mishra, J.S. Speck, S.A. Ringel, J. Appl.
Phys. 98 (2005) 053704.
[139] A. Armstrong, A.R. Arehart, B. Moran, S.P. DenBaars, U.K. Mishra, J.S. Speck, S.A.
Ringel, Appl. Phys. Lett. 84 (2004) 374.
[140] C.R. Abernathy, J.D. Mackenzie, S.J. Pearton, W.S. Hobson, Appl. Phys. Lett. 66
(1995) 1969.
[141] T. Hikosaka, N. Koide, Y. Honda, M. Yamaguchi, N. Sawaki, Phys. Stat. Sol. C 3
(2006) 14251428.
[142] U. Kohler, M.L. ubbers, J. Mimkes, D.J. As, Physica B 308310 (2001) 126129.
[143] A. Zadon, E. Tschumak, J.W. Gerlachb, K. Lischkaa, D.J. As, J. Cryst. Growth 323
(2011) 8890.
[144] H. Tang, J.B. Webb, J.A. Bardwell, S. Raymond, J. Salzman, C. Uzan-Saguy, Appl.
Phys. Lett. 78 (2001) 757.
[145] A.Y. Polyakov, N.B. Smirnov, E.A. Kozhukhova, A.V. Osinsky, S.J. Pearton, J. Vac.
Sci. Technol. B 31 (2013) 051208.
[146] C. Poblenz, P. Waltereit, S. Rajan, S. Heikman, U.K. Mishra, J.S. Speck, J. Vac. Sci.
Technol. B 22 (2004) 1145.
[147] S. Haffouz, H. Tang, J.A. Bardwell, E.M. Hsu, J.B. Webb, S. Rolfe, Solid-State
Electron. 49 (2005) 802807.
[148] S.A. Chevtchenko, E. Cho, F. Brunner, E. Bahat-Treidel, J. Wur, Appl. Phys. Lett.
100 (2012) 223502.
[149] P. Lu, R. Collazo, R.F. Dalmau, G. Durkaya, N. Dietz, B. Raghothamachar, M.
Dudley, Z. Sitar, J. Cryst. Growth 312 (2009) 58.
[150] M. Bickermann, B.M. Epelbaum, O. Filip, P. Heimann, S. Nagata, A. Winnacker,
Phys. Stat. Sol. C 7 (2010) 21.
[151] H. Helava, T. Chemekova, O. Avdeev, E. Mokhov, S. Nagalyuk, Y. Makarov, M.
Ramm, Phys. Stat. Sol. C 7 (2010) 2115.
[152] I. Nagai, T. Kato, T. Miura, H. Kamata, K. Naoe, K. Sanada, H. Okumura, J. Cryst.
Growth 312 (2010) 2699.
[153] B.N. Bryant, D.S. Kamber, F. Wu, S. Nakamura, J.S. Speck, Phys. Stat. Sol. C 8 (2011)
1463.
[154] T. Nomura, K. Okumura, H. Miyake, K. Hiramatsu, O. Eryu, Y. Yamada, J. Cryst.
Growth 350 (2012) 69.
[155] T. Nagashima, Y. Kubota, T. Kinoshita, Y. Kumagai, J. Xie, R. Collazo, H. Murakami,
H. Okamoto, A. Koukitu, Z. Sitar, Appl. Phys. Express 5 (2012) 125501.
[156] R. Collazo, J. Xie, B.E. Gaddy, Z. Bryan, R. Kirste, M. Hoffmann, R. Dalmau, B.
Moody, Y. Kumagai, T. Nagashima, Y. Kubota, T. Kinoshita, A. Koukitu, D.L. Irving,
Z. Sitar, Appl. Phys. Lett. 100 (2012) 191914.
[157] B. Monemar, O. Lagerstedt, J. Appl. Phys. 50 (1979) 64806491.
[158] J. Baur, K. Maier, M. Kunzer, U. Kaufmann, J. Schneider, H. Amano, I. Akasaki, T.
Detchprohm, K. Hiramatsu, Appl. Phys. Lett. 64 (1999) 857859.
[159] S. Heikman, S. Keller, S.P. DenBaars, U.K. Mishra, Appl. Phys. Lett. 81 (2002) 439
441.
[160] R.P. Vaudo, X.P. Xu, A. Salant, J. Malcarne, G.R. Brandes, Phys. Stat. Sol. A 200
(2003) 1821.
[161] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, S.J. Pearton, Appl. Phys. Lett. 83
(2003) 33143316.
[162] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, S.J. Pearton, J. Vac. Sci. Technol. B 22
(2004) 120.

54

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

[163] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.A. Shlensky, K. McGuire, E. Harley,
L.E. McNeil, R. Khanna, S.J. Pearton, J.M. Zavada, Phys. Stat. Sol. C 2 (2005) 2476
2479.
[164] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.A. Shlensky, S.J. Pearton, J. Appl.
Phys. 95 (2004) 55915595.
[165] M.1 Rudzinski, V. Desmaris, P.A. van Hal, J.L. Weyher, P.R. Hageman, K. Dynefors,
T.C. Rodle, H.F.F. Jos, 5H. Zirath, P.K. Larsen, Phys. Stat. Sol. C 3 (2006) 22312236.
[166] Z. Bougria, M. Azize, P. Lorenzini, M. Laught, H. Haas, Phys. Stat. Sol. A 202 (2005)
536.
[167] T. Aggerstam, M. Sjodin, S. Lourdudoss, Phys. Stat. Sol. C 3 (2006) 2373.
[168] S. Heikman, S. Keller, T. Mates, S.P. DenBaars, U.K. Mishra, J. Cryst. Growth 248
(2003) 513.
[169] M. Rudzinski, V. Desmaris, P.A. van Hal, J.L. Weyher, P.R. Hageman, K. Dynefors,
T.C. Rodle, H.F.F. Jos, H. Zirath, P.K. Larsen, Phys. Stat. Sol. C 3 (2006) 2231.
[170] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, T.G. Yugova, E.A.
Petrova, A.M. Dabiran, A.M. Wowchak, A.V. Osinski, P.P. Chow, S.J. Pearton,
K.D. Shcherbatchev, V.T. Bublik, J. Electrochem. Soc. 154 (2007) H749H754.
[171] F. Lipski, Semi-insulating GaN by Fe-Doping in Hydride Vapor Phase Epitaxy
Using a Solid Iron Source, Annual Report, Institute of Optoelectronics, Ulm
University, 2010, pp. 6370.
[172] R. Heitz, P. Maxim, L. Eckey, P. Thurian, A. Hoffmann, I. Broser, K. Pressel, B.K.
Meyer, Phys. Rev. B 55 (1997) 43824387.
[173] E. Malguth, A. Hoffmann, M.R. Phillips, Phys. Stat. Sol. B 245 (2008) 455.
[174] E. Malguth, A. Hoffmann, W. Gehlhoff, O. Gelhausen, M.R. Phillips, X. Xu, Phys.
Rev. B 74 (2006) 165202.
[175] P. Zakrzewski, Boguslawski, Acta Phys. Pol. A 124 (2013) 898900.
[176] T. Aggerstam, A. Pinos, S. Marcinkevicius, M. Linnarsson, S. Lourdudoss, J.
Electron. Mater. 36 (2007) 16211624.
[177] P. Muret, J. Pernot, M. Azize, Z. Bougrioua, J. Appl. Phys. 102 (2007) 053701.
[178] Z.-Q. Fang, B. Clafn, D.C. Look, S. Elhamni, H.E. Smith, W.C. Mitchel, D. Hanser,
E.A. Preble, Phys. Stat. Sol. C 5 (2008) 15081511.
[179] J. Dashdorj, M.E. Zvanut, J.G. Harrison, K. Udwary, T. Paskova, J. Appl. Phys. 112
(2012) 013712.
[180] U. Sunay, J. Dashdorj, M.E. Zvanut, J.G. Harrison, J.H. Leach, K. Udwary, in: A.
Cavallini, S. Estreicher (Eds.), AIP Proceedings, Melville, New York, vol. 1583,
2013, pp. 297301.
[181] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, T.G. Yugova, A.V. Markov, A.M.
Dabiran, A.M. Wowchak, B. Cui, J. Xie, A.V. Osinsky, P.P. Chow, S.J. Pearton,
Appl. Phys. Lett. 92 (2008) 042110.
[182] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, A.M. Dabiran, A.M.
Wowchak, B. Cui, A.V. Osinski, P.P. Chow, S.J. Pearton, J. Appl. Phys. 104 (2008)
053702.
[183] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, T.G. Yugova, E.A.
Petrova, H. Amano, T. Kawashima, K.D. Scherbatchev, V.T. Bublik, J. Appl. Phys.
105 (2009) 063708.
[184] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, E.B. Yakimov, P.S.
Vergeles, In-Hwan Lee, Cheul Ro Lee, S.J. Pearton, J. Vac. Sci. Technol. B 26
(2008) 990994.
[185] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, H. Amano, S.J. Pearton, I.-H. Lee, Q.
Sun, J. Han, S.Y. Karpov, Appl. Phys. Lett. 98 (2011) 072104.
[186] A.Y. Polyakov, N.B. Smirnov, M.-W. Ha, C.-K. Hahn, E.A. Kozhukhova, A.V.
Govorkov, R.V. Ryzhuk, N.I. Kargin, H.-S. Cho, I.-H. Lee, J. Alloys Compd. 575
(2013) 1723.
[187] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, C.-K. Hahn, S.J. Pearton, J. Vac. Sci. Technol.
B 32 (2014) 050602.
[188] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, N.Y. Pashkova, J. Kim, F. Ren, M.E.
Overberg, G.T. Thaler, C.R. Abernathy, S.J. Pearton, R.G. Wilson, J. Appl. Phys. 92
(2002) 31303137.
[189] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, N.V. Pashkova, A.A. Shlensky, S.J.
Pearton, M.E. Overberg, C.R. Abernathy, J.M. Zavada, R.G. Wilson, J. Appl. Phys. 93
(2003) 53885396.
[190] A.Y. Polyakov, A.V. Govorkov, N.B. Smirnov, N.Y. Pashkova, G.T. Thaler, M.E.
Overberg, R. Frazier, C.R. Abernathy, S.J. Pearton, J. Kim, F. Ren, J. Appl. Phys. 92
(2002) 49894993.
[191] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, A.M. Dabiran, P.P.
Chow, A.M. Wowchak, S.J. Pearton, J. Appl. Phys. 107 (2010) 023708.
[192] D.C. Look, D.C. Reynolds, J.W. Hemsky, J.R. Sizelove, R.L. Jones, R.J. Molnar, Phys.
Rev. B 79 (1997) 2273.
[193] G.A. Umana-Membreno, J.M. Dell, T.P. Hessler, et al. Appl. Phys. Lett. 80 (2002)
4354.
[194] G.A. Umana-Membreno, J.M. Dell, G. Parish, B.D. Nener, L. Faraone, U.K. Mishra,
IEEE Trans. Electron Devices 50 (2003) 2326.
[195] L. Polenta, Z.-Q. Fang, D.C. Look, Appl. Phys. Lett. 76 (2000) 20862088.
[196] A. Castaldini, A. Cavallini, L. Polenta, J. Phys. C: Condens. Matter 12 (2000) 10161.
[197] S.A. Goodman, F.D. Auret, M.J. Legodi, B. Beaumont, P. Gibart, Appl. Phys. Lett. 78
(2001) 3815.
[198] M. Hayes, F.D. Auret, L. Wu, W.E. Meyer, J.M. Nel, M.J. Legodi, Physica B 340342
(2003) 421425.
[199] S.A. Goodman, F.D. Auret, F.K. Koschnick, J.-M. Spaeth, B. Beaumont, P. Gibart,
Mater. Sci. Eng. B 71 (2000) 100103.
[200] A.Y. Polyakov, N.B. Smirnov, C.H. Roh, C.K. Hahn, H.-S. Cho, I.-H. Lee, E.A.
Kozhukhova, A.V. Govorkov, R.V. Ryzhuk, N.I. Kargin, IEEE Trans. Nanotechnol.
13 (2014) 151159.
[201] H.K. Cho, F.A. Khan, I. Adesida, Z.-Q. Fang, D.C. Look, J. Phys. D: Appl. Phys. 41
(2008) 155314.

[202] J. Bourgoin, M. Lannoo, Point Defects in Semiconductors. II: Experimental


Aspects, Springer Verlag, Berlin, 1983 (chapter 6).
[203] A.Y. Polyakov, In-Hwan Lee, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, N.G.
Kolin, A.V. Korulin, V.M. Boiko, S.J. Pearton, J. Appl. Phys. 109 (2011) 123703.
[204] A.Y. Polyakov, D.-W. Jeon, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, E.B.
Yakimov, I.-H. Lee, J. Appl. Phys. 112 (2012) 073112.
[205] H.K. Cho, C.S. Kim, C.-H. Hong, J. Appl. Phys. 94 (2003) 1485.
[206] A.R. Arehart, A. Corrion, C. Poblenz, J.S. Speck, U.K. Mishra, S.A. Ringel, Appl. Phys.
Lett. 93 (2008) 112101.
[207] A.R. Arehart, A. Corrion, C. Poblenz, et al. Phys. Stat. Sol. C 5 (2008) 1750.
[208] D. Emiroglu, J.H. Evans-Freeman, M.J. Kappers, C. McAleese, C.J. Humphreys,
Phys. Stat. Sol. C 5 (2008) 14821484.
[209] A.Y. Polyakov, A.S. Usikov, B. Theys, N.B. Smirnov, A.V. Govorkov, F. Jomard, N.M.
Shmidt, W.V. Lundin, Solid-State Electron. 44 (2000) 19711983.
[210] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, Q. Sun, Y. Zhang, C.D.
Yerino, T.-S. Ko, I.-H. Lee, J. Han, Mater. Sci. Eng. B 166 (2010) 220224.
[211] A.Y. Polyakov, L.-W. Jang, D.-S. Jo, I.-H. Lee, N.B. Smirnov, A.V. Govorkov, E.A.
Kozhukhova, K.H. Baik, S.-M. Hwang, J. Appl. Phys. 111 (2012) 033103.
[212] T. Ito, Y. Terada, T. Egawa, Mater. Res. Soc. Symp. 1068 (2008) C06C09.
[213] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, S.J. Pearton,
Thin Solid Films 516 (2008) 20352040.
[214] A.Y. Polyakov, N.B. Smirnov, A.S. Usikov, A.V. Govorkov, B.V. Pushniy, Solid-State
Electron. 42 (1998) 1959.
[215] A. Hierro, D. Kwon, S.A. Ringel, et al. Appl. Phys. Lett. 76 (2000) 3064.
[216] A. Hierro, S.A. Ringel, M. Hansen, J.S. Speck, U.K. Mishra, S.P. DenBaars, Appl.
Phys. Lett. 77 (2000) 1499.
[217] L. Polenta, A. Castaldini, A. Cavallini, J. Appl. Phys. 102 (2007) 063702.
[218] G.A. Umana-Membreno, G. Parish, N. Fichtenbaum, S. Keller, U.K. Mishra, B.D.
Nener, J. Electron. Mater. 37 (2008) 569.
[219] Z. Zhang, A.R. Arehart, E. Cinkilic, J. Chen, E.X. Zhang, D.M. Fleetwood, R.D.
Schrimpf, B. McSkimming, J.S. Speck, S.A. Ringel, Appl. Phys. Lett. 103 (2013)
042102.
[220] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, E.B. Yakimov, P.S.
Vergeles, N.G. Kolin, D.I. Merkurisov, V.M. Boiko, I.-H. Lee, C.-R. Lee, S.J. Pearton,
J. Electron. Mater. 36 (2007) 13201325.
[221] Z.-Q. Fang, D.C. Look, D.H. Kim, I. Adesida, Appl. Phys. Lett. 87 (2005) 182115.
[222] Z.-Q. Fang, B. Clain, D.C. Look, D.S. Green, R. Vetury, J. Appl. Phys. 108 (2010)
063706.
[223] P. Hacke, T. Detchprohm, K. Hiramatsu, N. Sawaki, J. Appl. Phys. 76 (1994) 304.
[224] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, I.-H. Lee, J.H. Baek, N.G. Kolin, V.M.
Boiko, D.I. Merkurisov, S.J. Pearton, J. Electrochem. Soc. 155 (2008) H31H35.
[225] A.Y. Polyakov, D.-W. Jeon, I.-H. Lee, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, E.B. Yakimov, J. Appl. Phys. 113 (2013) (2013) 083712.
[226] D. Pons, J. Appl. Phys. 55 (1984) 3644.
[227] Y. Tokuda, Y. Matuoka, K. Yoshida, H. Ueda, O. Ishiguro, N. Soejima, T. Kachi, Phys.
Stat. Sol. C 4 (2007) 2568.
[228] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, N.G. Kolin, D.I. Merkurisov, V.M. Boiko, K.D. Shcherbatchev, V.T. Bublik, M.I. Voronova, I.-H. Lee, C.R.
Lee, J. Appl. Phys. 100 (2006) 093715093719.
[229] Y.S. Puzyrev, T. Roy, E.X. Zhang, D.M. Fleetwood, R.D. Schrimpf, S.T. Pantelides,
IEEE Trans. Nucl. Sci. 58 (2011) 29182924.
[230] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, Z.-Q. Fang, D.C. Look, R.J. Molnar, A.V.
Osinsky, J. Appl. Phys. 91 (2002) 65806584.
[231] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, Z.-Q. Fang, D.C. Look, S.S. Park, J.H.
Han, J. Appl. Phys. 92 (2002) 52415247.
[232] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, J.M. Redwing, Solid-State Electron.
42 (1998) 831.
[233] I.-H. Lee, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, N.G.
Kolin, V.M. Boiko, A.V. Korulin, S.J. Pearton, J. Electrochem. Soc. 158 (2011)
H866H871.
[234] A.Y. Polyakov, N.B. Smirnov, E.B. Yakimov, A.S. Usikov, H. Helava, K.D. Shcherbachev, A.V. Govorkov, Yu N. Makarov, I.-H. Lee, J. Alloys Compd. 617 (2014)
200206.
[235] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, T.G. Yugova, H. Cox, A.S. Usikov, H.
Helava, Y. Makarov, J. Appl. Phys. 115 (2014) 183706.
[236] J. Oila, V. Ranki, J. Kivioja, K. Saarinen, P. Hautojarvi, J. Likonen, J.M. Baranowski,
K. Pakula, T. Suski, M. Leszczynski, I. Grzegory, Phys. Rev. B 63 (2001) 045205.
[237] M. Andrew, B.S. Armstrong, Investigation of deep level defects in GaN:C, GaN:Mg
and pseudomorphic AlGaN/GaN lms, (Ph.D. dissertation), The Ohio State
University, 2006.
[238] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, I.-H. Lee, S.J. Pearton, N.G. Kolin, I.L.
Gazizov, V.M. Zalyetin, Proc. SPIE 79451F (2011).
[239] M.A. Reshchikov, A. Usikov, H. Helava, Yu Makarov, Appl. Phys. Lett. 104 (2014)
032103.
[240] W. Gotz, N.M. Johnson, M.D. Bremser, R.F. Davis, Appl. Phys. Lett. 69 (1996) 2379.
[241] M.J. Legodi, S.S. Hullavarad, S.A. Goodman, M. Hayes, F.D. Auret, Physica B 308
310 (2001) 1189.
[242] Y.S. Park, C.J. Park, C.M. Park, J.H. Na, J.S. Oh, I.T. Yoon, H.Y. Cho, T.W. Kang, J.-E.
Oh, Appl. Phys. Lett. 86 (2005) 152109.
[243] J. Osaka, Y. Ohno, S. Kishimoto, K. Maezawa, T. Mizutani, Appl. Phys. Lett. 87
(2005) 222112.
[244] A.R. Arehart, A.A. Allerman, S.A. Ringel, J. Appl. Phys. 109 (2011) 114506.
[245] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, N.V. Pashkova, A.A. Shlensky, K.H.
Baik, S.J. Pearton, B. Luo, F. Ren, J.M. Zavada, J. Vac. Sci. Technol. B 22 (2004) 77
81.

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156
[246] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, N.V. Pashkova, S.J. Pearton, J.M.
Zavada, R.G. Wilson, J. Vac. Sci. Technol. B 21 (2004) 25002505.
[247] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, N.G. Kolin, V.M. Boiko,
D.I. Merkurisov, S.J. Pearton, J. Vac. Sci. Technol. B 24 (2006) 10941097.
[248] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, T.G. Yugova, K.D. Scherbatchev, O.A.
Avdeev, T.Y. Chemekova, E.N. Mokhov, S.S. Nagalyuk, H. Helava, Y.N. Makarov,
Physica B 404 (2009) 49394941.
[249] I. Vurgaftman, J.R. Meyer, J. Appl. Phys. 94 (2003) 3675.
[250] O. Ambacher, J. Smart, J.R. Shealy, N.G. Weimann, K. Chu, M. Murphy, W.J. Schaff,
L.F. Eastman, R. Dimitrov, L. Wittmer, M. Stutzmann, W. Rieger, J. Hilsenbeck, J.
Appl. Phys. 85 (1999) 32223233.
[251] L. Hsu, W. Walukiewicz, J. Appl. Phys. 89 (2001) 17831789.
[252] J.P. Ibbetson, P.T. Fini, K.D. Ness, S.P. DenBaars, J.S. Speck, U.K. Mishra, Appl. Phys.
Lett. 77 (2000) 250.
[253] I.P. Smorchkova, C.R. Elsas, J.P. Ibbetson, R. Vetury, B. Heying, P. Fini, E. Haus, S.P.
DenBaars, J.S. Speck, U.K. Mishra, J. Appl. Phys. 86 (1999) 4520.
[254] B. Jun, S. Subramanian, IEEE Trans. Nucl. Sci 49 (2002) 3222.
[255] S.Y. Karpov, in: Z.C. Feng (Ed.), III-Nitride Devices and Nanoengineering, Imperial
College Press, London, 2008, pp. 367398 (chapter 13).
[256] http://www.str-soft.com/products/FETIS/.
[257] X.Z. Dang, C.D. Wang, E.T. Yu, K.S. Boutros, J.M. Redwing, Appl. Phys. Lett. 72
(1998) 27452747.
[258] B.K. Li, K.J. Chen, K.M. Lau, W.K. Ge, J.N. Wang, Phys. Stat. Sol. C 5 (2008) 1892
1894.
[259] Y. Cai, Y. Zhou, K.J. Chen, K.M. Lau, IEEE Electron. Device Lett. 26 (2005) 435437.
[260] Y. Cai, Y.G. Zhou, K.M. Lau, K.J. Chen, IEEE Trans. Electron Devices 53 (2006) 2207.
[261] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, S.J. Pearton, F. Ren,
S.Y. Karpov, K.D. Shcherbachev, N.G. Kolin, W. Lim, J. Vac. Sci. Technol. B 30
(2012) 041209.
[262] O. Mitrofanov, M. Manfra, Superlattices Microstruct. 34 (2003) 3353.
[263] G. Koley, V. Tilak, L.F. Eastman, M.G. Spencer, IEEE Trans. Electron Devices 50
(2003) 886893.
[264] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, A.M. Dabiran, A.M.
Wowchak, A.V. Osinski, P.P. Chow, S.J. Pearton, Appl. Phys. Lett. 91 (2007)
232116.
[265] C.F. Lo, F. Ren, S.J. Pearton, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, I.A.
Belogorokhov, A.I. Belogorokhov, V.Y. Reznik, J.W. Johnson, J. Vac. Sci. Technol. B
29 (2011) 042201.
[266] C.F. Lo, L. Liu, T.S. Kang, F. Ren, O. Laboutin, Y. Cao, J.W. Johnson, A.Y. Polyakov,
N.B. Smirnov, A.V. Govorkov, I.A. Belogorokhov, A.I. Belogorokhov, S.J. Pearton, J.
Vac. Sci. Technol. B 30 (2012) 011205.
[267] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, S.J. Pearton, F. Ren,
L. Lu, J.W. Johnson, R.V. Ryzhuk, N.I. Kargin, J. Vac. Sci. Technol. B 31 (2013)
011211.
[268] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, S.J. Pearton, F. Ren,
J.W. Johnson, J. Vac. Sci. Technol. B 30 (2012) 061207.
[269] Y.-S. Hwang, L. Liu, F. Ren, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A.
Kozhukhova, N.G. Kolin, V.M. Boiko, S.S. Vereyovkin, V.S. Ermakov, C.-F. Lo, O.
Laboutin, Y. Cao, J.W. Johnson, N.I. Kargin, R.V. Ryzhuk, S.J. Pearton, J. Vac. Sci.
Technol. B 31 (2013) 022206.
[270] T. Imada, D. Piedra, T. Kikkawa, T. Palacios, Phys. Stat. Sol. A 211 (2014) 779783.
[271] W. Chikhaoui, J.M. Bluet, C. Bru-Chevallier, C. Dua, R. Aubry, Phys. Stat. Sol. C 7
(2010) 9295.
[272] M. Chargeddine, M. Gassami, H. Mosbachi, C. Caquire, M.A. Zaidi, H. Maaref, J.
Mod. Phys. 2 (2011) 12291234.
[273] H. Mosbahi, M. Gassoumi, H. Mejri, M.A. Zaidi, C. Caquire, H. Maaref, J. Electron.
Dev. 15 (2012) 12251231.
[274] Y. Nakano, Y. Irokawa, M. Takeguchi, Appl. Phys. Express 1 (2008) 091101.
[275] E.J. Miller, X.Z. Dang, H.H. Wieder, P.M. Asbeck, E.T. Yu, G.J. Sullivan, J.M.
Redwing, J. Appl. Phys. 87 (2000) 80708073.
[276] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, E.A. Kozhukhova, S.J. Pearton, F. Ren,
S.Y. Karpov, K.D. Shcherbatchev, W. Lim, ECS J. Solid State Sci. Technol. 1 (2012)
P152P156.
[277] D. Runton, B. Trabert, J. Shealy, R. Vetury, IEEE Microwave Mag. 14 (2013) 82.
[278] M.A. Khan, M.S. Shur, Q.C. Chen, J.N. Kuznia, Electron. Lett. 30 (1994) 2175.
[279] P.B. Klein, J.A. Freitas Jr., S.C. Binari, A.E. Wickenden, Appl. Phys. Lett. 75 (1999)
4014.
[280] W. Kruppa, S.C. Binari, K. Doverspike, Electron. Lett. 31 (1995) 1951.
[281] E. Kohn, I. Daumiller, P. Schmid, N.X. Nguyen, C.N. Nguyen, Electron. Lett. 35
(1999) 1022.
[282] I. Daumiller, D. Theron, C. Gaquiere, A. Vescan, R. Dietrich, A. Wieszt, H. Leier, R.
Vetury, U.K. Mishra, I.P. Smorchkova, S. Keller, N.X. Nguyen, C.N. Nguyen, E.
Kohn, IEEE Electron Device Lett. 22 (2001) 62.
[283] D.V. Kuksenkov, H. Temkin, R. Gaska, J.W. Yang, IEEE Electron Device Lett. 19
(1998) 222.
[284] S.L. Rumyantsev, N. Pala, M.S. Shur, E. Borovitskaya, A.P. Dmitriev, M.E. Levinshtein, R. Gaska, M.A. Khan, J. Yang, X. Hu, G. Simin, IEEE Trans. Electron Devices 48
(2001) 530.
[285] S.C. Binari, P.B. Klein, T.E. Kaizor, Proc. IEEE 90 (2002) 1048.
[286] O. Mitrofanov, M. Manfra, Appl. Phys. Lett. 84 (2004) 422424.
[287] A. Koudymov, C.X. Wang, V. Adivarahan, J. Yang, G. Simin, M.A. Khan, IEEE
Electron Device Lett. 28 (2007) 5.
[288] A. Koudymov, M. Shur, Int. J. High Speed Electron. Syst. 18 (2008) 935.
[289] R. Vetury, Q. Zhang, S. Keller, U.K. Mishra, IEEE Trans. Electron Devices 48 (2001)
560.

55

[290] G. Simin, A. Koudymov, A. Tarakji, X. Hu, J. Yang, M.A. Khan, M.S. Shur, R. Gaska,
Appl. Phys. Lett. 79 (2001) 2651.
[291] Y.F. Wu, A. Saxler, M. Moore, R.P. Smith, S. Sheppard, P.M. Chavarkar, T. Wisleder,
U.K. Mishra, P. Parikh, IEEE Electron Device Lett. 25 (2004) 117119.
[292] T.M.M. Higashiwaki, T. Matsui, Appl. Phys. Express 1 (2008) 021103.
[293] B.M. Green, K.K. Chu, E.M. Chumbes, J.A. Smart, J.R. Shealy, L.F. Eastman, IEEE
Electron Device Lett. 21 (2000) 268270.
[294] H. Leier, A. Vescan, R. Dietrich, A. Wieszt, H.H. Sledzik, IEICE Trans. Electron. E84C (2001) 14421447.
[295] S. Arulkumaran, T. Egawa, H. Ishikawa, T. Jimbo, Y. Sano, Phys. Lett. 84 (2004)
613615.
[296] A.P. Edwards, J.A. Mittereder, S.C. Binari, D.S. Katzer, D.F. Storm, J.A. Roussos, IEEE
Electron Device Lett. 26 (2005) 225227.
[297] B. Luo, J.W. Johnson, J. Kim, R.M. Mehandru, F. Ren, B.P. Gila, A.H. Onstine, C.R.
Abernathy, S.J. Pearton, A.G. Baca, R.D. Briggs, R.J. Shul, C. Monier, J. Han, Appl.
Phys. Lett. 80 (2002) 16611663.
[298] C. Liu, E.F. Chor, L.S. Tan, Appl. Phys. Lett. 88 (2006) 173504.
[299] A.V. Vertiatchikh, L.F. Eastman, W.J. Schaff, T. Prunty, Electron Lett. 38 (2002)
388.
[300] A.P. Zhang, L.B. Rowland, E.B. Kaminsky, V. Tilak, J.C. Grande, J. Teetsov, A.
Vertiatchikh, L.F. Eastman, J. Electron. Mater. 32 (2003) 388394.
[301] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, V.N. Danilin, T.A. Zhukova, B. Luo, F.
Ren, B.P. Gila, A.H. Onstine, C.R. Abernathy, S.J. Pearton, Appl. Phys. Lett. 83
(2003) 2608.
[302] Y.-H. Chen, K. Zhang, M.-Y. Cao, S.-L. Zhao, J.-C. Zhang, X.-H. Ma, Y. Hao, Appl.
Phys. Lett. 104 (2014) 153509.
[303] J.J.M. Law, E.T. Yu, G. Koblmuller, F. Wu, J.S. Speck, Appl. Phys. Lett. 96 (2010)
102111.
[304] J. Joh, J.A. del Alamo, IEEE Trans. Electron. Devices 58 (2011) 132140.
[305] M. Meneghini, D. Bisi, D. Marcon, S. Stoffels, M. Van Hove, T.-L. Wu, S. Decoutere,
G. Meneghesso, E. Zanoni, Appl. Phys. Lett. 104 (2014) 143505.
[306] W.C. Liao, Y.L. Chen, C.H. Chen, J.I. Chyi, Y.M. Hsin, Appl. Phys. Lett. 104 (2014)
033503.
[307] T. Mizutani, T. Okino, K. Kawada, Y. Ohno, S. Kishimoto, K. Mzezawa, Phys. Stat.
Sol. A 200 (2003) 195198.
[308] A.R. Arehart, A. Sasikumar, S. Rajan, G.D. Via, B. Poling, B. Winningham, E.R.
Heller, D. Brown, Y. Pei, F. Recht, U.K. Mishra, S.A. Ringel, Solid-State Electron. 80
(2013) 1922.
[309] A. Sasikumar, A.R. Arehart, S. Martin-Horcajo, M.F. Romero, Y. Pei, D. Brown, F.
Recht, M.A. di Forte-Poisson, F. Calle, M.J. Tadjer, S. Keller, S.P. DenBaars, U.K.
Mishra, S.A. Ringel, Appl. Phys. Lett. 103 (2013) 033509.
[310] A.R. Arehart, A.C. Malonis, C. Poblenz, Y. Pei, J. Speck, U. Mishra, S. Ringel, Phys.
Stat. Sol. C 8 (2011) 2242.
[311] A.R. Arehart, A. Sasikumar, S. Rajan, G.D. Via, B. Poling, B. Winningham, E.R.
Heller, D. Brown, Y. Pei, F. Recht, U.K. Mishra, S.A. Ringel, Solid-State Electron. 80
(2013) 19.
[312] D.W. Cardwell, A. Sasikumar, A.R. Arehart, S.W. Kaun, J. Lu, S. Keller, J.S. Speck,
U.K. Mishra, S.A. Ringel, J.P. Pelz, Appl. Phys. Lett. 102 (2013) 193509.
[313] M. Tapajna, R.J.T. Simms, Y. Pei, U.K. Mishra, M. Kuball, IEEE Electron Device Lett.
31 (2010) 662664.
[314] M. Montes Bajo, H. Sun, M.J. Uren, M. Kuball, Appl. Phys. Lett. 104 (2014)
223506.
[315] M.A. Khan, M.S. Shur, Q.C. Chen, J.N. Kuznia, Electron. Lett. 30 (1994) 21752176.
[316] P.B. Klein, S.C. Binari, J.A. Freitas Jr., A.E. Wickenden, J. Appl. Phys. 88 (2000)
2843.
[317] S.C. Binari, K. Ikossi, J.R. Roussos, W. Kruppa, D. Park, H. Dietrich, D.D. Koleske,
A.E. Wickenden, R.L. Henry, IEEE Trans. Electron Devices 48 (2001) 565.
[318] P.B. Klein, S.C. Binari, K. Ikossi-Anastasiou, A.E. Wickenden, D.D. Koleske, R.L.
Henry, D.S. Katzer, Electron. Lett. 37 (2001) 661.
[319] M. Tapajna, S.W. Kaun, M.H. Wong, F. Gao, T. Palacios, U.K. Mishra, J.S. Speck, M.
Kuball, Appl. Phys. Lett. 99 (2011) 223501.
[320] F. Gao, B. Lu, L. Li, S. Kaun, J.S. Speck, C.V. Thompson, T. Palacios, Appl. Phys. Lett.
99 (2011) 223506.
[321] A. Hinoki, J. Kikawa, T. Yamada, T. Tsuchiya, S. Kamiya, M. Kurouchi, K. Kosaka, T.
Araki, A. Suzuki, Y. Nanishi, Appl. Phys. Express 1 (2008) 011103.
[322] M.A. Lampert, R.B. Shilling, in: R.K. Willardson, A.C. Beer (Eds.), Semiconductors
and Semimetals,, Vol. 6, Academic Press, New York, 1970, p. 1.
[323] A.Y. Polyakov, S.J. Pearton, P. Frenzer, F. Ren, L. Liu, J. Kim, J. Mater. Chem. C 1
(2013) 877887.
[324] B.D. White, M. Bataiev, S.H. Goss, X. Hu, A. Karmarkar, D.M. Fleetwood, R.D.
Schrimpf, W.J. Schaff, L.J. Brillson, IEEE Trans. Nucl. Sci. 50 (2003) 19341941.
[325] A.P. Karmarkar, B. Jun, D.M. Fleetwood, R.D. Schrimpf, R.A. Weller, B.D. White, L.J.
Brillson, U.K. Mishra, IEEE Trans. Nucl. Sci. 51 (2004) 38013806.
[326] X. Hu, A.P. Karmarkar, B. Jun, D.M. Fleetwood, R.D. Schrimpf, R.D. Geil, R.A.
Weller, B.D. White, M. Bataiev, L.J. Brillson, U.K. Mishra, IEEE Trans. Nucl. Sci. 50
(2003) 17911796.
[327] B. Luo, F. Ren, K.K. Allums, B.P. Gila, A.H. Onstine, C.R. Abernathy, S.J. Pearton, R.
Dwivedi, T.N. Fogarty, R. Wilkins, R.C. Fitch, J.K. Gillespie, T.J. Jenkins, R. Dettmer,
J. Sewell, G.D. Via, A. Crespo, A.G. Baca, R.J. Shu, Solid-State Electron. 47 (2003)
10151020.
[328] B. Luo, J.W. Johnson, F. Ren, K.K. Allums, C.R. Abernathy, S.J. Pearton, R. Dwivedi,
T.N. Fogarty, R. Wilkins, A.M. Dabiran, A.M. Wowchack, C.J. Polley, P.P. Chow,
A.G. Baca, Appl. Phys. Lett. 79 (2001) 21962198.
[329] H.-Y. Kim, J. Ahn, J. Kim, S.P. Yun, J.S. Lee, J. Ceram. Process. Res. 9 (2008) 155
157.

56

A.Y. Polyakov, I.-H. Lee / Materials Science and Engineering R 94 (2015) 156

[330] L. Liu, C.-F. Lo, Y. Xi, Y. Wang, F. Ren, S.J. Pearton, H.-Y. Kim, J. Kim, R.C. Fitch, D.s.E.
Walker Jr., K.D. Chabak, J.K. Gillespie, S.E. Tetlak, G.D. Via, A. Crespo, J. Vac. Sci.
Technol. B 31 (2013) 022201.
[331] X. Hu, B.K. Choi, H.J. Barnaby, D.M. Fleetwood, R.D. Schrimpf, S. Lee, S. ShojahArdalan, R. Wilkins, U.K. Mishra, R.W. Dettmer, IEEE Trans. Nucl. Sci. 51 (2004)
293297.
[332] B. Luo, J.W. Johnson, F. Ren, K.K. Allums, C.R. Abernathy, S.J. Pearton, A.M.
Dabiran, A.M. Wowchack, C.J. Polley, P.P. Chow, D. Schoenfeld, A.G. Baca, Appl.
Phys. Lett. 80 (2002) 604606.
[333] T. Roy, Y.S. Puzyrev, B.R. Tuttle, D.M. Fleetwood, R.D. Schrimpf, D.F. Brown, U.K.
Mishra, S.T. Pantelides, Appl. Phys. Lett. 96 (2010) 133503.
[334] T. Roy, E.X. Zhang, Y.S. Puzyrev, D.M. Fleetwood, R.D. Schrimpf, B.K. Choi, A.B.
Hmelo, S.T. Pantelides, IEEE Trans. Nucl. Sci. 57 (2010) 30603065.
[335] A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, A.V. Markov, S.J. Pearton, N.G. Kolin,
D.I. Merkurisov, V.M. Boiko, J. Appl. Phys. 98 (2005) 033529.
[336] B. Luo, J. Kim, F. Ren, J.K. Gillespie, R.C. Fitch, J. Sewell, R. Dettmer, G.D. Via, A.
Crespo, T.J. Jenkins, B.P. Gila, A.H. Onstine, K.K. Allums, C.R. Abernathy, S.J.
Pearton, R. Dwivedi, T.N. Fogarty, R. Wilkins, Appl. Phys. Lett. 82 (2003) 1428.
[337] A. Stocco, S. Gerardin, D. Bisi, S. Dalcande, F. Rampazzo, M. Meneghini, G.
Meneghesso, J. Gruneputt, B. Lambert, E. Zanoni, Microelectron. Reliab. 54
(2014) 22132216.
[338] E.E. Patrick, M. Choudhury, F. Ren, S.J. Pearton, M. Law, ECS J. Solid State Sci.
Technol. 4 (2015) Q21Q25.
[339] G. Wenping, C. Chi, D. Huantao, H. Yue, Z. Jincheng, W. Chong, F. Qian, M.
Xiaohua, J. Semicond. 30 (2009) 44002.
[340] E.J. Katz, C.-H. Lin, J. Qiu, Z. Zhang, U.K. Mishra, L. Cao, L.J. Brillson, J. Appl. Phys.
115 (2014) 123705.
[341] E. Kioupakis, Q. Yan, C.G. Van de Walle, Appl. Phys. Lett. 101 (2012) 231107.
[342] D.S. Meyaard, Q. Shan, Q. Dai, J. Cho, E. Fred Schubert, M.-H. Kim, C. Sone, Appl.
Phys. Lett. 99 (2011) 1112.
[343] K. Okamoto, Y. Kawakami, IEEE J. Sel. Top. Quantum Electron 15 (2009) 1199
1209.
[344] T. Deguchi, K. Sekiguchi, A. Nakamura, T. Sota, R. Matsuo, S. Chichibu, S.
Nakamura, Jpn. J. Appl. Phys. 2 38 (1999) L914.
[345] T. Kinoshita, K. Hironaka, T. Obata, T. Nagashima, R. l Dalmau, R. Schlesser, B.
Moody, J. Xie, S.-I. Inoue, Y. Kumagai, A. Koukitu, Z. Sitar, Appl. Phys. Express 5
(2012) 122101.
[346] E.F. Schubert, Light-Emitting Diodes, 2nd ed., Cambridge University Press,
London, 2006.
[347] www.str-soft.com/products/SILENSE.
[348] A.Y. Polyakov, A.V. Govorkov, N.B. Smirnov, A.V. Markov, I.-H. Lee, S. Jin-Woo Ju,
Y. Karpov, N.M. Shmidt, S.J. Pearton, J. Appl. Phys. 105 (2009) 123708.
[349] O. Soltanovich, E. Yakimov, Phys. Stat. Sol. C 10 (2013) 338341.
[350] D.V. Lang, M.B. Panish, F. Capasso, J. Alam, R.A. Hamm, A.M. Sergent, W.T. Tsang,
Appl. Phys. Lett. 50 (1987) 736.
[351] D.V. Lang, A.M. Sergent, M.B. Panish, H. Temkin, Appl. Phys. Lett. 49 (1986)
812.
[352] A. Polyakov, A. Govorkov, N. Smirnov, A. Markov, I.-H. Lee, H.-K. Ahn, S. Karpov,
S.J. Pearton, Phys. Stat. Sol. A 207 (2010) 13831385.
[353] J.S. Im, H. Kollmer, O. Gfrorer, J. Off, F. Scholz, A. Hangleiter, MRS Internet J.
Nitride Semicond. Res. 4S1 (1999) G6.20.
[354] E.B. Yakimov, P.S. Vergeles, A.Y. Polyakov, N.B. Smirnov, A.V. Govorkov, I.-H. Lee,
C.R. Lee, S.J. Pearton, Appl. Phys. Lett. 90 (2007) 152114.
[355] S. Nakamura, in: S.J. Pearton (Ed.), GaN and Related Materials II, Gordon and
Breach Science, The Netherlands, 2000, pp. 146.
[356] A. Armstrong, T.A. Henry, D.D. Koleske, M.H. Crawford, K.R. Westlake, S.R. Lee,
Appl. Phys. Lett. 101 (2012) 162102.
[357] A. Armstrong, T.A. Henry, D.D. Koleske, M.H. Crawford, S.R. Lee, Opt. Express 20
(2012) A812.

[358] E. Gur, Z. Zhang, S. Krishnamoorthy, S. Rajan, S.A. Ringel, Appl. Phys. Lett. 99
(2011) 092109.
[359] S.D. Brotherton, Solid-State Electron. 19 (1976) 341.
[360] M. Meneghini, M. la Grassa, S. Vaccari, B. Galler, R. Zeisel, P. Drechsel, B. Hahn, G.
Meneghesso, E. Zanoni, Appl. Phys. Lett. 104 (2014) 113505.
[361] D. Schiavon, M. Binder, M. Peter, B. Galler, P. Drechsel, F. Scholz, Phys. Stat. Sol. B
250 (2013) 283290.
[362] G. Venturi, A. Castaldini, A. Cavallini, M. Meneghini, E. Zanoni, D. Zhu, C.
Humphreys, Appl. Phys. Lett. 104 (2014) 211102.
[363] A.Y. Polyakov, J.-H. Yun, A.S. Usikov, E.B. Yakimov, N.B. Smirnov, K.D. Shcherbachev, H. Helava, Y.N. Makarov, S.Y. Kurin, N.M. Shmidt, O.I. Rabinovich, S.I.
Didenko, S.A. Tarelkin, B.P. Papchenko, I.-H. Lee, Thin Solid Films (2015).
[364] E.B. Yakimov, S.S.S.S. Borisov, S.I. Zaitsev, Semiconductors 41 (2007) 411413.
[365] E.A. Menkovich, S.A. Tarasov, I.A. Lamkin, A.V. Solomonov, S.Y. Kurin, A.A.
Antipov, I.S. Barash, A.D. Roenkov, A.S. Usikov, H.I.Y. Helava, N. Makarov, J.
Phys.: Conf. Ser. 541 (2014) 012054.
[366] M. Kuroda, C. Sasaoka, A. Kimura, A. Usui, Y. Mochizuki, J. Cryst. Growth 189/190
(1998) 551.
[367] M. Hansen, P. Fini, M. Craven, B. Heying, J.S. Speck, S.P. DenBaars, J. Cryst. Growth
234 (2002) 623.
[368] J.S. Hwang, A. Gokarna, Y.-H. Cho, J.K. Son, S.N. Lee, T. Sakong, H.S. Paek, O.H.
Nam, Y. Park, Appl. Phys. Lett. 90 (2007) 131908.
[369] O.H. Nam, K.H. Ha, J.S. Kwak, S.N. Lee, K.K. Choi, T.H. Chang, S.H. Chae, W.S. Lee,
Y.J. Sung, H.S. Paek, J.H. Chae, T. Sakong, J.K. Son, H.Y. Ryu, Y.H. Kim, Y. Park, Phys.
Stat. Sol. A 201 (2004) 27172720.
[370] M. Meneghini, N. Trivellin, K. Orita, S. Takigawa, T. Tanaka, D. Ueda, G. Meneghesso, E. Zanoni, Appl. Phys. Lett. 97 (2010) 263501.
[371] L. Marona, P. Wisniewski, P. Prystawko, I. Grzegory, T. Suski, S. Porowski, P.
Perlin, R. Czernecki, M. Leszczynski, Appl. Phys. Lett. 88 (2006) 201111.
[372] M. Meneghini, C. de Santi, N. Trivellin, K. Orita, S. Takigawa, T. Tanaka, D. Ueda, G.
Meneghesso, E. Zanoni, Appl. Phys. Lett. 99 (2011) 093506.
[373] M. Pavesi, M. Manfredi, G. Salviati, N. Armani, F. Rossi, G. Meneghesso, S. Levada,
E. Zanoni, S. Du, I. Eliashevich, Appl. Phys. Lett. 84 (2004) 3403.
[374] M. Rossetti, T.M. Smeeton, W.-S. Tan, M. Kauer, S.E. Hooper, J. Heffernan, H. Xiu,
C.J. Humphreys, Appl. Phys. Lett. 92 (2008) 151110.
[375] N. Shmidt, A. Usikov, E. Shabunina, A. Chernyakov, A. Sakharov, S. Kurin, A.
Antipov, I. Barash, A. Roenkov, H. Helava, Y. Makarov, Phys. Stat. Sol. C (2015),
http://dx.doi.org/10.1002/pssc.201400172.
[376] E.I. Shabunina, N.S. Averkiev, A.E. Chernyakov, M.E. Levinshtein, P.V. Petrov, N.M.
Shmidt, Phys. Stat. Sol. C 10 (2013) 335.
[377] N. Shmidt, A. Greshnov, A. Chernyakov, M. Levinshtein, A. Zakgeim, E. Shabunina,
Phys. Stat. Sol. C 10 (2013) 332.
[378] A.E. Chernyakov, M.E. Levinshtein, P.V. Petrov, N.M. Shmidt, E.I. Shabunina, A.L.
Zakgeim, Microelectron. Reliab. 52 (2012) 2180.
[379] X.D. Chen, Y. Huang, S. Fung, C.D. Beling, C.C. Ling, J.K. Sheu, M.L. Lee, G.C. Chi, S.J.
Chang, Appl. Phys. Lett. 82 (2003) 36713673.
[380] Z. Gong, M. Gaevski, V. Adivarahan, W. Sun, M. Shatalov, M. Asif Khan, Appl. Phys.
Lett. 88 (2006) 121106.
[381] A. Pinos, S. Marcinkevicius, J. Yang, Y. Bilenko, M. Shatalov, R. Gaska, M.S. Shur,
Appl. Phys. Lett. 95 (2009) 181914.
[382] S. Sawyer, S.L. Rumyantsev, M.S. Shur, Solid-State Electron. 52 (2008) 968.
[383] M. Meneghini, M. Pavesi, N. Trivellin, R. Gaska, E. Zanoni, G. Meneghesso, IEEE
Trans. Device Mater. Reliab. 8 (2008) 248.
[384] C.G. Moe, M.L. Reed, G.A. Garrett, A.V. Sampath, T. Alexander, H. Shen, M.
Wraback, Y. Bilenko, M. Shatalov, J. Yang, W. Sun, J. Deng, R. Gaska, Appl. Phys.
Lett. 96 (2010) 213512.
[385] A. Pinos, S. Marcinkevicius, M.S. Shur, J. Appl. Phys. 109 (2011) 103108.

Vous aimerez peut-être aussi