JAP Tutorial HEMT Rev10 D
JAP Tutorial HEMT Rev10 D
1
    University of Padova, Department of Information Engineering, via Gradenigo 6/B, 35131
Padova, Italy
2
    IEMN (Institute of Electronics, Microelectronics and Nanotechnology), Avenue Poincaré, 59650
Villeneuve d’Ascq, France
3
    École Polytechnique Fédérale de Lausanne, ELD 012 (Bâtiment ELD), CH-1015 Lausanne
4
    University of Modena and Reggio Emilia, Dipartimento di Scienze e Metodi dell'Ingegneria, via
Amendola 2 - Pad. Morselli, 42122 Reggio Emilia, Italy
5
    University of Modena and Reggio Emilia, Dipartimento di Ingegneria "Enzo Ferrari", Via P.
Vivarelli, 10, 41125 Modena, Italy
Abstract:
Over the last decade, gallium nitride has emerged as an excellent material for the fabrication of
power devices. Among the semiconductors for which power devices are already available on the
market, GaN has the widest energy gap, the largest critical field, the highest saturation velocity,
thus representing an excellent material for the fabrication of high speed/high voltage components.
For high power/high voltage operation, vertical device architectures are being proposed and
investigated, and 3-dimensional structures – fin-shaped, trench-structured, nanowire-based – are
demonstrating a great potential. Contrary to silicon, GaN is a relatively young material: trapping
and degradation processes must be understood and described in detail, with the aim of optimizing
device stability and reliability.
This tutorial paper describes the physics, technology and reliability of GaN-based power devices:
in the first part of the article, starting from a discussion of the main properties of the material, the
characteristics of lateral and vertical GaN transistors are discussed in detail, to provide guidance
in this complex and interesting field. The second part of the paper focuses on trapping and
reliability aspects: the physical origin of traps in GaN, and the main degradation mechanisms are
discussed in detail. The wide set of referenced papers and the insight on the most relevant aspects
gives the reader a comprehensive overview on present and next-generation GaN electronics.
1 Table of Contents
2     Introduction ........................................................................................................................................... 6
8.8.5 Interface trap characterization by means of C-V and G-V measurements ........................ 130
11 Acknowledgments......................................................................................................................... 168
Over the past decade, gallium nitride has emerged as an excellent material for the fabrication of
power semiconductor devices. Thanks to the unique properties of GaN, diodes and transistors
based on this material have excellent performance, compared to their silicon counterparts, and are
expected to find wide application in the next-generation power converters. Owing to the flexibility
and the energy efficiency of GaN-based power converters, the interest towards this technology is
rapidly growing: the aim of this tutorial is to review the most relevant physical properties, the
operating principles, the fabrication parameters, and the stability/reliability issues of GaN-based
power transistors. For introductory purposes, we start summarizing the physical reasons why GaN
transistors achieve a much better performance than the corresponding silicon devices, to help the
reader understanding the unique advantages of this technology.
The properties of GaN devices allow the fabrication of high-efficiency (near or above 99 %) [1]–
[5], kW-range power converters. Such converters can have switching frequencies above 1 MHz
[6], [7], and – through proper design, integration and/or hybrid GaN/CMOS manufacturing –
frequencies as high as 40-75 MHz can be reached [8], [9]. High frequency operation permits to
substantially reduce the size and weight of inductors and capacitors, thus resulting in a compact
converter design. Further innovation will come from the design of monolithically integrated all-
GaN integrated circuits: specific platforms, such as GaN on silicon-on-insulator (SOI) can be used
for the fabrication of fully integrated power converters, containing smart control, pulse width
modulation (PWM) circuitry, dead time control and half bridge [10], [11]. Such solutions, that can
be tailored for switching in the 1-10 MHz range, can reach very short turn on/off times, which are
considerably smaller than in discrete gate drivers [11]. The availability of fast, small, efficient and
light-weight power converters can be particularly beneficial in the fields of portable/consumer
electronics, automotive, and avionics.
GaN is a wide-bandgap semiconductor, and has an energy-gap of 3.4 eV [12]. This allows GaN
devices to be operated at extremely high temperatures, thus substantially increasing the maximum
power density that can be dissipated on a device, or permitting the use of light and small heat sinks.
Over the last decades, several report on high temperature and stable operation of GaN HEMTs
have been published. Temperatures above 400-500 °C [13]–[15] have been reached and, for
selected InAlN/GaN devices, up to 900 °C [15]. Operation at high temperature is a first,
substantial, advantage of GaN devices, compared to silicon-based Metal-Oxide-Semiconductor
Field Effect Transistors (MOSFETs), that are typically rated for maximum operating temperature
of 125-150 °C.
A second advantage of GaN arises from its high breakdown field (3.3 MV/cm [12]), which is 11
times higher than that of silicon (0.3 MV/cm). The direct consequence of such high critical field
is that for withstanding a given voltage, a layer of GaN can be 11 times thinner than its silicon
counterpart, with consequent beneficial impact on resistivity. As a consequence, the use of GaN
switches can substantially reduce the resistive losses in switching mode power supplies (SMPSs).
A third aspect to be considered is the high mobility of the channel: as will be discussed in the
following sections, GaN transistors are typically heterostructure devices. A high mobility (up to
2000 cm/Vs [12]) channel can be obtained through the formation of a 2-dimensional electron gas
(2DEG) at the heterointerface between the AlGaN barrier and the GaN channel layer. Such high
mobility, along with the large saturation velocity (2.5x107 cm/s), further contributes to reduce the
resistivity of the devices. FETs based on AlGaN/GaN heterostructures are usually referred to as
high electron mobility transistors (HEMTs), or heterojunction field effect transistors (HFETs).
At present, GaN devices are commercially available, and several products have been proposed, in
three main voltage ranges: a) low/mid-voltage (VDS,max<200 V) devices find application in dc-dc
power converters, motor drives, wireless power transfer, LiDAR and pulsed power applications,
solar micro-inverters, class-D audio amplifiers, robotics, and synchronous rectification. Such
devices can have on-resistances below 2 m (for drain currents up to 90 A) [16], or up to 100-200
m (for operating currents in the range 0.5-5 A), depending on the final application [17], [18]. b)
high voltage (VDS,max up to 650 V), that find application in telecommunication servers, industrial
converters, photovoltaic inverters, servo motor control [19], lighting applications, power adapters,
converters for consumer electronics [20], class D amplifiers [21], datacenter SMPS [22], [23]. c)
devices with ultra-high voltage (VDS,max above 1 kV). At present, no kV-range transistor based on
GaN is commercially available. The commercial transistors with highest voltage rating have a
maximum voltage of 900 V, and are expected to find application in data communication systems,
industrial application, motor control, and photovoltaic inverters. As will be discussed in the article,
several research papers demonstrated the feasibility of GaN transistors with breakdown voltages
above 1 kV [24]–[26], and proposed possible fabrication processes to target this voltage range.
kV-range GaN transistors will compete with SiC-based devices, in the industrial, automotive and
photovoltaic application environments.
As will be described in detail in the paper, current GaN devices have typically a lateral layout.
Several key aspects related to device design, fabrication and performance must be discussed in
detail, to understand how the performance of the transistors can be optimized through careful
device design: this will be done extensively in the following sections of this paper. For introductory
purposes, we remind here that in a HEMT current flows between drain and source through a 2-
dimensional electron gas (2DEG), which is formed at the heterojunction between an AlGaN barrier
and a GaN layer. Figure 1 (a) reports the schematic structure of a GaN-based HEMT, showing the
main layers that constitute the structure. Power GaN devices are typically grown on a silicon
substrate, to minimize cost and maximize yield. Growing GaN on a silicon substrate is particularly
complicated, due to: a) the large mismatch of the in-plane thermal expansion coefficient (2.6x10-
6
    K-1 for Si and 5.59x10-6 K-1 for GaN [27]), that may lead to cracking of the GaN layer during the
cooling-phase after the epitaxial growth; b) the large lattice mismatch (around 16 % for Si(111))
[28], [29], that may result in the propagation of dislocations through the gallium nitride epitaxial
layers, with consequent defect generation. A careful optimization of the buffer is needed in order
to limit the propagation of such defects towards the 2DEG region; Figure 1 (b) reports a cross
sectional SEM image of the epitaxial layers of a GaN-based transistor. As can be noticed, the use
of a step-graded buffer in combination with an AlN nucleation layer is used to release the strain
and prevent the formation/propagation of dislocations.
The lateral structure described in Figure 1 may have some limitations, when extremely high
breakdown voltages and/or power densities are targeted. First, in a lateral transistor the breakdown
voltage scales with the gate-drain spacing. Thus, devices with a high breakdown voltage can be
fabricated, but will be more resistive and will use a wider semiconductor area, thus resulting in a
higher device cost. Second, the density of electrons in the 2DEG can be strongly influenced by
surface charges: for this reason, the performance of the final devices is strongly dependent on the
process and backend. To solve the limitation of GaN lateral devices, vertical device structures are
currently being explored and investigated, in line with what has been done with silicon and silicon
carbide components. In a vertical device (see a schematic structure in Figure 1 (c)), current flows
through the bulk of the material, thus allowing high current and power densities. The density of
electrons in the channel is modulated through a metal-oxide-semiconductor (MOS) stack, and a p-
type body is usually employed to shift the threshold voltage towards more positive values. The
breakdown voltage of a vertical GaN transistor depends on the thickness of the lightly-doped drift
region, and not on the size and area of the device as in a lateral transistor. Vertical GaN devices
represent the latest development in GaN technology, and the reader will find interest in the related
concepts and applications, described in this paper.
As for every technology, there are some physical processes that may limit the performance and the
reliability of GaN devices. The on-resistance of a GaN transistors (and thus the density of electrons
in the 2DEG) strongly depends on the intrinsic (spontaneous and piezoelectric) polarization
charges of GaN, as well as on the presence of charges trapped at surface states (e.g. in the
passivation layer or at the interface between the passivation layer and the AlGaN barrier) or in
buffer states. For this reason, it is of fundamental importance to know and manage the surface-
and buffer-related trapping phenomena that may limit the dynamic performance of GaN
transistors, leading to a recoverable increase in on-resistance (dynamic-Ron problem). Trapping
in the epitaxial layers and/or at the gate-stack may result in positive- or negative-bias threshold
instability (PBTI or NBTI), and the related processes must be investigated and understood to be
able to fabricate fast and reliable devices.
Finally, GaN-based transistors are operated at field, temperature, and frequency levels which are
unimaginable for conventional silicon devices. Electric fields can be in excess of 3 MV/cm, and
channel temperatures can be above 300 °C during operation, if lightweight heat dissipators are
used. Such conditions may favor sudden or time-dependent breakdown phenomena, leading to the
failure of the devices. Furthermore, operation at high frequencies may exacerbate the degradation
processes related to hard switching events. For this reason, it is of utmost importance to understand
the degradation processes of GaN power devices, and to identify ways and strategies for improving
the robustness of the components.
This tutorial paper presents a detailed overview on the physics, performance and reliability of
GaN-based power devices. In the Section 3-5, the properties and physical parameters of gallium
nitride are discussed, to help the reader understanding the unique advantages offered by GaN
compared to other semiconductors. The main figures of merit (FOM) for high speed (Johnson
FOM) and high power (Baliga FOM) devices are also introduced. Finally, the properties of
AlGaN/GaN heterostructures, and the related band diagrams are described in detail, and first-order
formulas for the calculation of sheet electron charge and threshold voltage are introduced. In
Section 6, the properties, structure and characteristics of lateral GaN devices are discussed.
Specific details is given to the various approaches for normally-off operation, to the main device
parameters, and to the optimization of the buffer. A perspective on AlN-based devices and on
possible strategies to increase the breakdown voltage is given, also by discussing devices with
local substrate removal. Section 7 deals with GaN vertical devices. First, the advantages of vertical
GaN transistors are described. Attention is then given to the choice of the substrate (GaN-on-GaN
vs GaN-on-Si) for vertical device manufacturing. Then, the various vertical device architectures
are discussed and compared, in terms of performance and structural parameters. In Section 8, the
stability of GaN devices is analyzed in detail. Specific focus is given to the role of surface traps,
barrier traps and buffer traps in modifying the main device parameters, to present a clear overview
of the topic. A complete overview of the dominant defects and deep levels in GaN is given in
Section 8.1, to provide an exhaustive view of the problem. Finally, Section 9 describes the most
relevant degradation processes that can limit the lifetime of GaN-based transistors. Specific
attention will be towards the degradation mechanisms induced by exposure to off-state stress,
semi-on state regime, and on-state degradation (with focus on gate reliability for p-GaN and
insulated-gate devices).
Through a pedagogical approach, this paper helps the reader understanding the advantages of GaN
technology, and getting familiar with the main performance, design, and reliability aspects.
                                                                                n- drift region
                 Buffer layer(s)
             AlN nucleation layer
                 Si-substrate                                              n+ GaN and substrate
                                                                                        D
Figure 1: (a) schematic representation of the structure of a lateral GaN high-electron mobility
transistor (HEMT). The main layers constituting the structure are shown, as well as the three
contacts of source (S), gate (G) and drain (D). (b) Cross-sectional SEM images of the epi-structure
of GaN HEMT on silicon. Reprinted from "Strain Analysis of GaN HEMTs on (111) Silicon with
Two Transitional AlxGa1−xN Layers", Y. Cai et al., https://doi.org/10.3390/ma11101968, licensed
under CC BY 4.0 (http://dx.doi.org/10.3390/ma11101968) [30]. (c) schematic representation of a
vertical GaN trench-MOSFET
Gallium nitride, along with its InGaN and AlGaN alloys, represents an excellent material for both
optoelectronics and electronics. In the early GaN era, the research efforts on gallium nitride have
been driven by the need of fabricating high-efficiency short-wavelength (blue/violet) LEDs. GaN
has a direct bandgap of 3.4 eV, thus being ideal for manufacturing ultraviolet optoelectronic
devices. In addition, the energy gap of III-N alloys can be tuned between the 0.7 eV of InN and
the 6.2 eV of AlN thus, in principle, allowing fabrication of LEDs with ultraviolet (UVA, UVB,
UVC), visible and infrared emission.
Contrary to other semiconductors, like InP or GaAs, III-N semiconductors typically have a
wurtzite crystal, with its characteristic hexagonal shape (see Figure 2). It can be easily understood
that this lattice arrangement does not have an inversion plane perpendicular to the c-axis (0001)
and, for this reason, the surfaces have either atoms from group III (In, Ga, Al), or nitrogen atoms
[31]. The nature of the surfaces has a fundamental importance, since it determines the polarity of
the polarization charges, as will be discussed in the following.
Figure 2: (a) Hexagonal unit cell and (b) atomic structure of Ga- and N-polar GaN. The arrows
represent the direction of the spontaneous polarization dipole, P, in the GaN crystal. From "Recent
progress in metal-organic chemical vapor deposition of (0001) N-polar group-III nitrides", S.
Keller et al., Semiconductor Science and Technology, Volume 29, Number 11, 113001, August
2014, DOI: 10.1088/0268-1242/29/11/113001, IOP Publishing. Reproduced with permission. All
rights reserved. (https://iopscience.iop.org/article/10.1088/0268-1242/29/11/113001/pdf) [32].
Table 1 reports the main parameters of gallium nitride, as compared with other semiconductor
materials, including silicon, gallium arsenide, silicon carbide, aluminum nitride, diamond and
gallium oxide. The materials are ordered with increasing energy gap EG, from left to right.
Excluding the three semiconductors for which commercial devices are not available (gallium
oxide, diamond, and aluminum nitride), GaN is the semiconductor with the largest energy gap, the
largest critical field, and the highest saturation velocity. As a consequence, it is an ideal candidate
for the fabrication of power semiconductor devices, capable of operating at high temperature and
voltage levels.
     Breakdown field (MV/cm)                        Research-grade                                            Si
                                    (a)             devices
                                                                                                            10     (b)
                               10                                                     -Ga2O3                                GaAs
                                                                                                             1
                                        Commercially                                                        0.1
                                        available devices
0.01
1 Diamond GaN
                                                                                                   AlN              4H-SiC
                                    1                       2        3   4 5 6 7
                                                 Energy gap (eV)
Figure 3 (a) reports the relation between breakdown field and energy gap for the semiconductor
materials in Table 1. As can be noticed, breakdown field has a power-law dependence on the
energy gap, in the form 𝐸𝑐𝑟𝑖𝑡 ∝ 𝐸𝐺2.3 . This dependence is consistent with previous reports in the
literature [33], and demonstrates the great advantage of using wide bandgap semiconductors for
fabricating electron devices with high breakdown voltage.
Since a single parameter does not fully describe the properties of a material, semiconductors are
typically compared by using figures of merit: the most commonly used are the Johnson FOM, for
high-speed devices, and the Baliga FOM, for high power devices. With regard to the first, it is
defined as the product of the maximum voltage and the maximum transit frequency, for a given
value of the drain-source spacing (see details in Ref. [34]), i.e. as:
                                                                                                𝐸𝑐𝑟𝑖𝑡 𝑣𝑠
                                                                𝐽𝑜ℎ𝑛𝑠𝑜𝑛 𝐹𝑂𝑀 = 𝑓𝑇 𝑉𝐷𝑆,𝑚𝑎𝑥 =
                                                                                                  2𝜋
The Johnson figure of merit indicates that, in general, devices with high breakdown voltage are
typically slower than devices with lower voltage rating.
Figure 3 (b) reports the values of the Johnson figure of merit for the semiconductors listed in Table
1; all values are normalized to that of GaN, to allow an easy comparison. As can be noticed, the
Johnson FOM of GaN is slightly higher than that of SiC (0.556), comparable to -Ga2O3 (0.978),
and is much larger than that of silicon (0.0379) and GaAs (0.0455). AlN and diamond are better
than GaN, but their Johnson FOMs are between 2.5 and 3, thus having the same order of magnitude
of GaN. While there is a substantial advantage by moving from Si/GaAs to GaN, the improvement
obtained by changing to AlN and diamond is only incremental, in terms of the Johnson FOM [34].
While the Johnson FOM allows to compare semiconductor materials for the fabrication of high-
speed devices, the Baliga FOM has been introduced to compare semiconductors for application in
power electronics. Considering a unilateral and abrupt (e.g. p+/n-) junction, as depicted in Figure
4 (a), the electric field profile at the n-side has a triangular shape (Figure 4 (b)), and – at the critical
electric field Ecrit – the space charge region has a width equal to
where ND is the donor density at the n-side of the diode, and 𝜖𝑟,𝐺𝑎𝑁 𝜖0 is the product between the
relative permittivity of GaN and the permittivity of vacuum . The breakdown voltage is therefore
                                                     1 2
                                         1             𝐸𝑐𝑟𝑖𝑡 𝜖𝑟,𝐺𝑎𝑁 𝜖0
                                  𝑉𝑏𝑟   = 𝐸𝑐𝑟𝑖𝑡 𝑊𝑑 = 2
                                         2                𝑞𝑁𝐷
The resistance RON of the n-type semiconductor region is proportional to 𝑊𝑑 /𝜇𝑁𝑑 , so we can
calculate [35]
                                   𝑊𝑑       𝑊𝑑2                 2
                                                             4𝑉𝑏𝑟
                           𝑅𝑂𝑁 ∝      ∝                ∝            3
                                   𝜇𝑁𝐷 𝜇𝜖𝑟,𝐺𝑎𝑁 𝜖0 𝐸𝑐𝑟𝑖𝑡 𝜇𝜖𝑟,𝐺𝑎𝑁 𝜖0 𝐸𝑐𝑟𝑖𝑡
The denominator of the last term is linked to the Baliga FOM [36], [37] (that was initially defined
as 𝜇𝜖𝐸𝐺3 ), and identifies the material parameters that help minimizing the conduction losses in
power transistors. This FOM is defined based on the assumption that power losses only originate
from the on-resistance of the FET. For this reason, it applies at relatively moderate frequencies,
where conduction losses are dominant [36]. For higher frequency devices, one would have to
consider also the contribution of switching losses. Figure 4 (c) reports the Baliga figure of merit
                  3
(calculated as 𝜇𝜖𝐸𝑐𝑟𝑖𝑡 ) for the same set of semiconductors in Table 1. All values are normalized to
GaN: the plot indicates that the Baliga FOM of GaN is substantially larger than those of silicon
and GaAs, higher than SiC, similar to the one of gallium oxide. Diamond and AlN (ultra wide-
bandgap semiconductors) have a much higher Baliga FOM, and can be considered as interesting
alternatives to further push the limits.
                                                     Research-grade                                                Si
                                     (a)             devices
                                                                                                                 10      (b)
                                10                                                            -Ga2O3                              GaAs
                                                                                                                  1
                                         Commercially                                                            0.1
                                         available devices
0.01
1 Diamond GaN
                                                                                                        AlN               4H-SiC
                                     1                       2          3   4 5 6 7
                                                   Energy gap (eV)
Figure 3: (a) dependence of breakdown field on energy gap for the semiconductor materials in
Table 1. (b) Johnson figure of merit for the same set of semiconductors. All values are normalized
to GaN, to allow an easy comparison. Data are taken from Table 1
                                                                            Si
                                                                         100
    p+                                              -Ga2O3               10
                                                                                              GaAs
                                                                           1
0.1
0.01
n- 0.001
Diamond GaN
AlN 4H-SiC
Figure 4: (a) schematic representation of a unilateral abrupt p+/n- junction; (b) approximated
                                                                                             3
electric field profile for the junction in (a); (c) Baliga figure of merit (calculated as 𝜇𝜖𝐸𝑐𝑟𝑖𝑡 ) for the
same set of semiconductors in Table 1. All values are normalized to GaN, to allow an easy
comparison. Data are taken from Table 1
The dependence of the bandgap of GaN on temperature follows the Varshni relation
                                                           𝛼 𝑇2
                                        𝐸𝐺 (𝑇) = 𝐸𝐺,0 −
                                                           𝑇+𝛽
For GaN, AlN and InN the related parameters are summarized in Table 2. As can be noticed, by
using alloys of GaN, AlN and InN it is possible to vary the bandgap of the alloy in a wide range,
from 0.7 eV to 6.2 eV. In most cases, GaN transistors are based on AlGaN/GaN heterostructures,
and AlN and AlGaN layers are used as nucleation and buffer layers respectively. InAlN devices
have also been investigated: lattice matched InAlN/GaN HEMTs allow an efficient down-scaling
of the transistor dimensions, thus allowing to reach high cut-off frequencies [38].
For ternary alloys, such as AlGaN and InGaN, the bandgap deviates from the Vegard’s rule, and
follows the empirical expression
where 𝐸𝐺 (𝐴𝑁) and 𝐸𝐺 (𝐵𝑁) are the bandgaps of the two materials (AN and BN), x is the molar
fraction of A, and b is a bowing parameter. For AlGaN, the material of interest for AlxGa(1-x)N/GaN
HEMTs, the relation has been determined as [39], [40]
Other bandstructure parameters of interest for GaN are the effective density of states in the
conduction and valence bands (𝑁𝐶 = 2.24 ∙ 1018 𝑐𝑚−3, 𝑁𝑉 = 4.56 ∙ 1019 𝑐𝑚−3), and the
effective masses of electrons and holes (𝑚𝑒 = 0.20 and 𝑚ℎ = 1.49) [34]
Table 2: bandgap and Varshni parameters for GaN, AlN and InN. For GaN, results data are taken
from [41], for AlN from [42], for InN from [43].
Contrary to conventional semiconductors, GaN has a unique advantage, that helps obtaining high
carrier densities even in absence of intrinsic doping: gallium nitride, in fact, is a polar material,
and exhibits strong polarization effects.
As already mentioned, the wurtzite crystal of III-N semiconductors, which are typically grown
epitaxially along the (0001) orientation, leads to the existence of polarization fields, that are both
spontaneous and piezoelectric. With zero external field, the total polarization 𝑃 is equal to the sum
of the spontaneous polarization 𝑃𝑠𝑝 and of the piezoelectric (or strain-induced) polarization 𝑃𝑝𝑧 .
Bernardini et al. [44] investigated the polarization in GaN layers along the (0001) axis. Nitrogen
has a higher electronegativity, compared to gallium. As a consequence, Ga and N atoms have
anionic (+) and cationic (-) characteristics, generating a spontaneous polarization 𝑃𝑠𝑝 along the
(0001) axis [35]. Wurtzite is the crystal arrangement with highest symmetry compatible with the
presence of spontaneous polarization [44]–[46]. The arrangement of the cation and anion
sublattices can lead to a relative movement from the ideal wurtzite position, favoring the
spontaneous polarization [47]. The orientation of spontaneous polarization is defined assuming
that the positive direction goes from the metal (Ga) to the nearest nitrogen atom, along the c-axis
[48]. Figure 5 depicts the crystal structure of GaN, and the sign and direction of the spontaneous
polarization.
The values of the spontaneous polarization in GaN, InN and AlN are reported in Table 3 for binary
semiconductors. In the case of interfaces between binary and ternary semiconductors, the
following expressions can be used (see also [49] and references therein):
Figure 5: crystal structure of GaN, showing the sign and direction of the spontaneous polarization
(adapted from Ref. [50])
Table 3: values of spontaneous polarization of III-N semiconductors (binary). Values in the first
row are from [44], values in second row are from [51].
The strain in the crystal, and the displacement of the anion sublattice with respect to the cation
sublattice, can lead to a piezoelectric polarization of the III-N semiconductors. A (simplified)
representation of the polarization is given in Figure 6, that reports a ball and stick diagram of the
bond (theahedral) between gallium and nitrogen. In this figure, the Ga-polar configuration is
represented. The electron cloud is closer to the nitrogen atoms, and this generates the polarization
vectors. If the tetrahedron is ideal, the in-plane and vertical polarization components cancel each
other [52]. When an in-plane tensile strain is applied (as shown in Figure 6, (a)) the polarization
generated by the triple bonds decreases, and this generates a net polarization along the (0001̅)
direction. On the contrary, when an in-plane compressive strain is applied (as shown in Figure 6,
(b)) the polarization generated by the triple bonds increases, and this generates a net polarization
along the (0001) direction.
To calculate the piezoelectric polarization, one needs to refer to the piezoelectric constants of the
materials under analysis. Details on the main parameters were given in [44], [48], [53], and a
comprehensive summary was presented in [49]; here, in Table 4, we report the values of
piezoelectric constants 𝑒𝑖𝑗 and lattice parameters 𝑎0 and 𝑐0 for GaN, InN and AlN.
As discussed in [48], the lattice structure of wurtzite semiconductors is defined by the length of
the hexagonal edge 𝑎0 , the height of the prism 𝑐0 , and a parameter 𝑢 that defines the length of the
bond parallel to the c-axis ([0001]) in units of 𝑐0 . To calculate the piezoelectric polarization 𝑃𝑝𝑧
along the c-axis, the key relation is
Here 𝜖𝑧 = (𝑐 − 𝑐0 )/𝑐0 represents the strain along the c-axis, while the in-plane strain 𝜖𝑥 = 𝜖𝑦 =
(𝑎 − 𝑎0 )/𝑎0 is assumed to be isotropic. 𝑎 and 𝑐 are the lattice constants of the strained layers, and
differ from 𝑎0 𝑎𝑛𝑑 𝑐0.
By considering that the lattice constants in a hexagonal AlGaN system are related according to
                                  𝑐 − 𝑐0       𝐶13
                                         = −2 ( ) (𝑎 − 𝑎0 )/𝑎0
                                    𝑐0         𝐶33
where 𝐶13 and 𝐶33 are elastic constants, the value of piezoelectric polarization along the c-axis can
be calculated as
                                                             𝑒33 𝐶13
                                𝑃𝑝𝑧 = 2(𝑎 − 𝑎0 )/𝑎0 (𝑒31 −           )
                                                              𝐶33
For an AlxGa1-xN/GaN stack, the material system of interest for GaN-based transistors, the values
of the elastic constants can be calculated from the following formulas (see [49], [54] for details)
and the variation of the lattice constant with the molar fraction is
Figure 6: schematic ball-and-stick configuration of a GaN tethrahedron with in-plane (a) tensile
and (b) compressive strain, showing a net polarization. A full description including also the case
of N-polar material can be found in [52].
     𝑒31 (𝐶 ∙ 𝑚−2 )       -0.49 [44]                -0.57 [44]              -0.6 [44], -0.58 [54]
      𝑒33 (𝐶 ∙ 𝑚−2 )      0.73 [44]                 0.97 [44]                 1.46 [44], 1.55 [54]
Table 4: values of piezoelectric constants 𝑒𝑖𝑗 and lattice parameters 𝑎0 and 𝑐0 for GaN, InN and
AlN
In a GaN layer, polarization charges are present on each unit cell. As schematically depicted in
Figure 7, the internal polarization charges cancel each other. Only the 𝑄𝜋 and −𝑄𝜋 at the N- and
Ga-faces remain, and form a charge dipole (Figure 7 (c)). Based on the numbers given in Table 3,
the spontaneous polarization charge in GaN has values in the range 1.8 − 2.1 ∙ 1013 𝑒 − /𝑐𝑚2 . In
absence of other charges, the polarization charge would lead to the presence of a dipole, resulting
in a fairly high electric field, in the range of MV/cm. Such dipole is screened through the formation
of a screening dipole (𝑄𝑠𝑐𝑟 ). In absence of the screening dipole, a non-physical situation would
form. As discussed in [47], a first hypothesis would be that the screening dipole originates from
ions from the atmosphere (𝐻 + , 𝑂𝐻 − ); however, the dipole is present also in material grown in
atmosphere free of counter ions, like during molecular beam epitaxy (MBE) growth.
A different interpretation can be given by considering the presence of donor states at the surface
of the GaN layer [55], [56], as schematically represented in Figure 8 (a). In absence of
compensating charge, due to the presence of the polarization dipole ±𝑄𝜋 , an electric field is present
in the GaN layer, and the band diagram shows a linear slope (Figure 8 (b)). If the GaN layer is
sufficiently thick, the donor states pins the Fermi level at the surface (at the donor level 𝐸𝐷𝐷 ), and
                      +
the screening charge 𝑁𝐷𝐷 at the surface is formed. The presence of a surface level 𝐸𝐷𝐷 has been
proved also experimentally [47]. Realistic GaN layers typically have a n-type conductivity, as
mentioned above. The ionized (bulk) donors also contribute to the overall charge balance. For a
thick GaN layer (see Figure 8 (d)), the bands are nearly flat (corresponding to negligible field),
apart from the surface region. The polarization charge −𝑄𝜋 at the surface is compensated by the
                                                +
positive charge of the ionized surface defects 𝑁𝐷𝐷 and by the total density of charges in the
depleted n-type GaN [47].
                                              ..
                                               ..
Figure 7: (a) schematic representation of a GaN lattice with ball and stick representation of the
bonds, and indication of the Ga- and N-faces, along the (0001) direction. (b) model for polarization
charge in a gallium nitride layer. (c) charge distribution at the Ga- and N-faces, showing the
polarization and screening dipoles (see also [47])
Figure 8: (a) schematic representation of a n-type GaN layer grown on a substrate. The (defective)
interface region is shown. (b) charge diagram showing the screening induced by surface donors.
(c) band diagram showing the surface donor level approaching the Fermi level at the surface of
                                                              +
GaN, thus leading to the generation of the screening charge 𝑞𝑁𝐷𝐷 . (d) band diagram for a thick
GaN (n-type), considering the presence of surface states and of donor charges (figure adapted from
[47])
The core of a GaN-based HEMT is the AlGaN/GaN heterostructure. Both the GaN and the AlGaN
layers are typically left undoped, to minimize electron scattering at impurities. Based on the
considerations above, undoped GaN has a weak n-type conductivity, with electron densities that
depend on the quality of the epitaxy, and that – even in the best case – are higher than 1015 cm-3
[47] (early GaN films were showing a high n-type conductivity, in the range of 1017-1018 cm-3 [57],
[58]). Optimizing the background electron density in lateral and vertical HEMTs requires a tuning
of the growth process, and the control of the residual impurities (such as carbon, see for instance
[59]).
Figure 9 shows the charge distribution, the electric field profile, and a schematic band diagram for
an AlGaN/GaN heterostructure used in HEMT technology. In the figure, 𝑡 is the thickness of the
AlGaN layer, while the 2DEG is supposed to be located at a position 𝑑, a few nanometers far from
the heterojunction, on the GaN side.
At the surface of the AlGaN layer, the total charge is determined by the sum of the charge of the
                                                                    +
surface donors and of the polarization charge of the AlGaN layer, 𝑞𝑁𝐷𝐷 − 𝑄𝜋(𝐴𝑙𝐺𝑎𝑁) . At the
AlGaN/GaN heterojunction, the total charge is given by the difference between the polarization
charges of AlGaN and GaN, i.e. 𝑄𝜋(𝐴𝑙𝐺𝑎𝑁) − 𝑄𝜋(𝐺𝑎𝑁) . For gallium polar material, this is a positive
number, since AlGaN has a higher polarization, compared to GaN. In the triangular potential well
formed near the heterojunction, on the GaN side, a 2-dimensional electron gas is formed. For
simplicity, we consider this sheet of charge to be located in 𝑥 = 𝑑 (position of the centroid of the
electron distribution).
The value of 𝑑 can be easily determined by solving the Schroedinger equation in the triangular
potential well formed at the AlGaN/GaN interface, that is schematically represented in Figure 10.
The electric field ℇ𝐺𝑎𝑁 is supposed to originate only from the charge in the 2DEG (𝑛𝑠 ), and is thus
equal to 𝑞𝑛𝑠 /(𝜖0 𝜖𝐺𝑎𝑁 ). The solution of the Schroedinger equation is approximated by the formula
[60]–[62]
                       1             2         2           1         2       2           2
                ℏ 3 3         3    3 3    ℏ 3 3    3    3 3 𝑞𝑛𝑠 3
         𝐸𝑛 ≈ ( ∗ ) ( 𝜋𝑞ℇ𝐺𝑎𝑁 ) (𝑛 + ) = ( ∗ ) ( 𝜋𝑞) (𝑛 + ) (        )
               2𝑚    2             4     2𝑚    2        4   𝜖0 𝜖𝐺𝑎𝑁
We now consider that the first sub-band 𝐸0 is dominant. The 2D density of states associated with
a single quantized level is [62]
                                                     𝑞𝑚∗
                                            𝐷𝐷𝑂𝑆   =
                                                     𝜋ℏ2
The electron concentration in the 2DEG can then be calculated starting from the 2D density of
states and from the position of the Fermi level, by using the Fermi-Dirac distribution as follows
[62]:
                                          𝑘𝑇               𝑞(𝐸𝐹 − 𝐸0 )
                            𝑛𝑠 = 𝐷𝐷𝑂𝑆 ∙      ∙ ln [1 + exp             ]
                                           𝑞                  𝑘𝑇
By simple calculations (see details in Ref. [47]), one can calculate the position 𝑑 of the 2DEG,
with respect to the heterointerface. Here the 2DEG is considered as an ideal 2D sheet of electrons
located at a distance 𝑑 from the interface, on the GaN side. For an AlGaN/GaN heterostructure, a
typical value of 𝑑 is 2 nm [47].
At the bottom of the GaN layer, at the interface with the substrate, a screening charge compensates
the (positive) component 𝑄𝜋(𝐺𝑎𝑁) . The reader should note that in more realistic structures other
layers (e.g. a C-doped layer, a step-graded or a superlattice-based buffer, an AlN nucleation layer)
are placed between the GaN channel layer and the substrate; these layers are not shown here for
simplicity.
At the surface of the AlGaN layer (if the layer is thick enough), the potential 𝑞𝜙𝑠 is pinned at the
surface donor level 𝐸𝐷𝐷 . The AlGaN layer is not doped, so – in absence of external bias – its
electric field depends only the sheet charge densities as follows:
                                               +
                                            [𝑞𝑁𝐷𝐷 − 𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁)]
                                 ℇ𝐴𝑙𝐺𝑎𝑁   =
                                                  𝜖0 𝜖𝐴𝑙𝐺𝑎𝑁
With regard to the electric field in the GaN, we suppose that it originates only from the charge in
the 2DEG, and neglect (for this analysis) the effect of the n-type donor charge in the GaN. A more
realistic solution can be obtained numerically, through technology computer-aided design (TCAD)
tools. Under this assumption, the electric field in the GaN is
                                                    𝑞𝑛𝑠
                                          ℇ𝐺𝑎𝑁 =
                                                   𝜖0 𝜖𝐺𝑎𝑁
                                 +
                              [𝑞𝑁𝐷𝐷 − 𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁)]    [𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁) − 𝑄𝜋 (𝐺𝑎𝑁) − 𝑞𝑛𝑠 ]
     𝑉𝐴𝑙𝐺𝑎𝑁   = −ℇ𝐴𝑙𝐺𝑎𝑁 𝑡 = −                     𝑡=                                𝑡
                                    𝜖0 𝜖𝐴𝑙𝐺𝑎𝑁                  𝜖0 𝜖𝐴𝑙𝐺𝑎𝑁
                                                        𝑞𝑛𝑠
                                    𝑉𝐺𝑎𝑁 = ℇ𝐺𝑎𝑁 𝑑 =            𝑑
                                                       𝜖0 𝜖𝐺𝑎𝑁
                                                 Δ𝐸𝑐
                                 𝜙𝑠 − 𝑉𝐴𝑙𝐺𝑎𝑁 −       + 𝑉𝐺𝑎𝑁 = 0
                                                  𝑞
By combining the equations above, by considering 𝜖 = 𝜖0 𝜖𝐺𝑎𝑁 ~𝜖0 𝜖𝐴𝑙𝐺𝑎𝑁 , the value of the sheet-
charge density 𝑛𝑠 can be calculated
                                                              𝛥𝐸
                            [𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁) − 𝑄𝜋 (𝐺𝑎𝑁)]𝑡 − 𝜖 (𝜙𝑠 − 𝑞 𝐶 )
                       𝑛𝑠 =
                                           𝑞(𝑡 + 𝑑)
When a potential 𝑉𝐺 is applied to the gate (through a suitable metal deposited on the AlGaN
barrier), the charge in the 2DEG can be modulated: a more positive voltage will fill the channel,
whereas moving to negative values of 𝑉𝐺 will lead to the depletion of the 2DEG.
                                                                     Δ𝐸
                             [𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁) − 𝑄𝜋 (𝐺𝑎𝑁)]𝑡 + 𝜖 [𝑉𝐺 − (𝜙𝑏 − 𝑞 𝐶 )]
                  𝑛𝑠 (𝑉𝐺 ) =
                                                𝑞(𝑡 + 𝑑)
where 𝜙𝑏 is the barrier at the metal/AlGaN interface.
Figure 11 (a) reports the value of the polarization charge 𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁) − 𝑄𝜋 (𝐺𝑎𝑁) at the
AlxGa1-xN/GaN interface as a function of the molar fraction 𝑥. The values have been calculated
based on the parameters given in Ref, [54] as:
                                                                  𝐶
                               𝑄𝜋,𝑠𝑝 (𝑥) = (−0.052𝑥 − 0.029) [       ]
                                                                  𝑚2
A typical range of interest for the mole fraction is between 0.2 and 0.4; such values are sufficiently
high to give a reasonable polarization charge (which is necessary for generating electrons in the
2DEG), and enough conduction band discontinuity at the AlGaN/GaN heterointerface (which is
necessary to ensure a good confinement of the 2DEG electrons). At the same time, the use of molar
fractions higher than 0.4 may be critical, since the thermal and lattice mismatch between AlGaN
and GaN may lead to high defect density and rough interfaces, that may limit the overall device
performance.
Figure 11 (b) reports the sheet charge density (𝑛𝑠 ) of the 2DEG for three AlxGa1-xN/GaN
interfaces, as a function of the Al mole fraction and of the thickness of the AlGaN layer. For the
calculation, a Schottky barrier equal to
was used, and the conduction band discontinuity at the AlxGa1-xN/GaN interface was calculated as
with
                   𝐸𝐺 (𝑥) = 𝑥 ∙ 6.13 𝑒𝑉 + (1 − 𝑥) ∙ 3.42 𝑒𝑉 − 𝑥(1 − 𝑥) ∙ 1 𝑒𝑉
As can be noticed, when the mole fraction 𝑥 is around 0.2-0.3, the sheet charge density is around
1013 𝑐𝑚−3. With increasing thickness of the AlGaN layer, the density of electrons in the 2DEG
increases, since the conduction band at the channel edge drops further below the Fermi level at the
AlGaN/GaN interface (see also the band diagram in Figure 9). As can be understood, AlGaN/GaN
HEMTs are intrinsically normally-on. At zero gate bias, the channel is formed, with a high electron
density. A first possible approach to change the threshold voltage (by keeping the same Al content
in the barrier) is to reduce the thickness of the AlGaN barrier. For sufficiently thin AlGaN barriers,
the electron density falls to zero, and the 2DEG vanishes. This effect has also been observed
experimentally, see for instance Ref. [55]. This is just one of the possible approaches to achieve
normally-off operation in HEMTs; an excessive thinning of the barrier, along with the related
etching process, may result in a significant increase in the leakage, and counter measures are
required to guarantee a good device performance. In the next sections, the various approaches for
normally-off operation are compared critically.
                                                      Solid: approximated profile
                                                      Dashed: realistic profile
Figure 9: charge distribution, electric field profile, and schematic band diagram of an AlGaN/GaN
heterostructure
                                                                             3
Figure 10: triangular potential well, similar to the one formed at the GaN-side of an AlGaN/GaN
interface. The energy varies linearly with field ℇ, starting from the value 𝑞𝑉(0) at the GaN-side
of an AlGaN/GaN interface. The energy levels and wavefunctions are schematically drawn (not to
scale). Figure is adapted from Ref [63].
                                                                                  t=10 nm
                                                                                  t=20 nm
                                                                                  t=30 nm
Figure 11: (a) polarization charge density at the AlGaN/GaN heterointerface (𝑄𝜋 (𝐴𝑙𝐺𝑎𝑁) −
𝑄𝜋 (𝐺𝑎𝑁)) calculated as described in [54]. (b) sheet charge density calculated for the 2DEG of an
AlGaN/GaN heterojunction as a function of molar fraction 𝑥 and thickness 𝑡 of the AlGaN layer
In a GaN-based transistor, a gate metal is placed on an AlGaN/GaN heterostructure; the resulting
schematic structure is shown in Figure 12 (a). For a power semiconductor device, the most critical
parameters are the on-resistance 𝑅𝑜𝑛 (that needs to be as low as possible, to minimize the resistive
losses in power converters) and the breakdown voltage (𝑉𝑏𝑟 , that must be sufficiently high to
ensure good reliability). In AlGaN/GaN HEMTs, these two parameters are strongly correlated:
carriers are generated by polarization, and the breakdown voltage scales with the distance between
the gate and drain (𝐿𝐺𝐷 ) [35]. In an first-order approximation, the on-resistance of the device is
the sum of the channel resistance (𝑅𝑐ℎ𝑎𝑛𝑛𝑒𝑙 , originated from 2DEG under the gate) and of the
drain-side access region (𝑅𝑑𝑟𝑎𝑖𝑛 ), according to the following equation [35]:
                                                     𝐿𝐺   1   𝐿𝐺𝐷   1
                       𝑅𝑜𝑛 = 𝑅𝑐ℎ𝑎𝑛𝑛𝑒𝑙 + 𝑅𝑑𝑟𝑎𝑖𝑛 =        ∙   +     ∙
                                                     𝑊𝐺 𝑞𝜇𝑛𝑠 𝑊𝐺 𝑞𝜇𝑛𝑠
It is worth noticing that in this calculation the contribution of the source and contact resistance is
neglected, with no loss of generality. The off-state voltage is supported by the gate-drain access
region; at the breakdown voltage, the relation between gate-drain spacing and the breakdown
voltage is 𝐿𝐺𝐷 = 𝑉𝑏𝑟 /𝐸𝑐𝑟𝑖𝑡 , and we can write:
                                                      1 1            𝑉𝑏𝑟
                        𝑅𝑜𝑛 = 𝑅𝑐ℎ𝑎𝑛𝑛𝑒𝑙 + 𝑅𝑑𝑟𝑎𝑖𝑛 =             (𝐿𝐺 +       )
                                                      𝑊𝐺 𝑞𝜇𝑛𝑠       𝐸𝑐𝑟𝑖𝑡
In high-voltage devices, the gate-drain spacing is typically longer than the gate length. Figure 6
(b) reports the variation of the on-resistance (multiplied by the gate width 𝑊𝑔 ) as a function of the
target breakdown voltage for devices having different gate lengths (0.25 µm and 1 µm). For
mobility and sheet charge density values of 1400 cm2/Vs and 1013 cm-2 were used. As can be
understood, for low-breakdown voltage devices (corresponding to devices with a short gate-drain
distance), the on-resistance is strongly determined by the gate length. Decreasing the gate length
from 1 µm to 0.25 µm can lead to a substantial reduction in the resistive losses. On the contrary,
for high breakdown voltage devices, the resistive contribution of the gate-drain access region
becomes relevant, and 𝑅𝑜𝑛 scales with the breakdown voltage. Figure 6 (b) also shows that the use
of GaN devices with nearly-ideal breakdown field (3.3 MV/cm, see Table 1) can lead to a
substantial improvement in on-resistance, compared to the more conservative case of 1 MV/cm
reported in previous publications [35]. Optimizing the breakdown field of the material is a key
step for minimizing the resistive losses in power semiconductor devices.
AlGaN
Figure 12: (a) schematic representation of the structure of a GaN HEMT, showing the parasitic
resistance of the channel (𝑅𝑐ℎ𝑎𝑛𝑛𝑒𝑙 ) and of the gate-drain access region (𝑅𝑑𝑟𝑎𝑖𝑛 ). (b) dependence
of the product 𝑅𝑜𝑛 ∙ 𝑊𝑔 on breakdown voltage, for devices with different gate length, and under
the hypothesis that the breakdown field is 1 MV/cm and 3.3 MV/cm
Besides resistive losses, also switching losses can play a relevant role in limiting the efficiency of
a switching mode power converter. One of the parameters used to quantify the switching losses
related to a specific transistor is the gate charge 𝑄𝐺 . For understanding the meaning of this
parameter, we consider the simple circuit in Figure 12 (a). Here the FET is supposed to be ideal,
and three capacitances 𝐶𝐺𝑆 , 𝐶𝐺𝐷 , 𝑎𝑛𝑑 𝐶𝐷𝑆 are added to model the parasitic capacitive components.
A fixed current is forced into the gate, and the measured gate voltage is plotted as a function of the
charge flowing towards the gate. Figure 12 (b) shows the schematic gate-charging curve, during a
turn-on event, i.e. when the device switches from the off-state (high drain voltage, zero current) to
the on-state (low-drain voltage, high current). At device turn-on, the gate-source voltage
𝑉𝐺𝑆 increases; when 𝑉𝐺𝑆 reaches the threshold voltage 𝑉𝑡ℎ , current starts flowing through the
device. At the same time, the gate-source capacitance 𝐶𝐺𝑆 is charged, until the 𝑉𝑀𝑖𝑙𝑙𝑒𝑟 voltage (and
the corresponding plateau) is reached. At this point, 𝐶𝐺𝑆 is completely charged, and the drain
current reaches the value fixed by the circuit. At the same time, 𝑉𝐺𝑆 becomes almost constant, and
the drive current starts charging the Miller capacitance 𝐶𝐺𝐷 . This process goes on until the
capacitance 𝐶𝐺𝐷 is fully charged. When both 𝐶𝐺𝑆 and 𝐶𝐺𝐷 are fully charged, the gate voltage starts
increasing again. Through this experimental procedure the charges 𝑄𝐺𝑆 and 𝑄𝐺𝐷 can be calculated.
The gate charge 𝑄𝐺𝑆 + 𝑄𝐺𝐷 is the minimum charge required to turn on the transistor [64], and is
thus representative of the switching losses.
A low value of 𝑄𝐺𝑆 + 𝑄𝐺𝐷 results in the device’s ability to achieve high commutation speed
(dV/dt), and in a substantial reduction in switching losses [65]. A low gate charge also results in a
reduced gate drive power for GaN devices, compared to silicon components [66]. Since, as stated
above, the dc losses depend on the 𝑅𝑜𝑛 𝑊𝐺 product, and the ac losses are proportional to 𝑄𝐺 /𝑊𝐺
[35], the 𝑅𝑜𝑛 ∙ 𝑄𝐺 product is an important parameter describing the switching efficiency of a given
device. In a recent paper [67], Chen et al. compared devices based on silicon, silicon carbide and
gallium nitride in terms of figures of merit. They showed that, for devices with comparable on-
resistance, the 𝑅𝑜𝑛 ∙ 𝑄𝐺 product can be around 3800 𝑚Ω ∙ nC for a silicon superjunction MOSFET,
in the range 1950 − 3480 𝑚Ω ∙ nC for SiC-based FETs, and around 290 − 300 𝑚Ω ∙ nC for a E-
mode GaN transistor. This result indicates that GaN E-mode HEMTs can contribute to a substantial
reduction in switching losses, compared to conventional semiconductor transistors.
The rapid evolution of wide bandgap semiconductors in the recent years has positioned lateral
GaN transistors as key enablers in the power device market. The interest in power applications has
undergone a remarkable shift due to the technological advantages of GaN HEMTs, which allow
for simultaneous high voltage, high current and low on-state resistance, resulting in high power
and high efficiency operation. In addition, the wide bandwidth provides a robust and reliable
technology capable of operating at high frequency and high temperature. This is why GaN-based
lateral power electronic devices are emerging as switching components for next generation high-
efficiency power converters. Furthermore, GaN-on-Silicon technology platform offers the best
cost figures for commercialization of these products, although technologically very challenging.
For instance, a complex GaN buffer for stress management and insulation purpose is required.
This paves the way for a growing number of applications in various fields, including consumer
electronics, transportation and energy, as well as several industrial, automotive and aerospace
applications, such as rectifiers and high-voltage converters. As can be seen in Figure 14, each
application uses a specific voltage range. In terms of device market, currently, the majority of GaN
components are designed for 600/650V applications and below. However, there are few companies
that offer devices for 900V applications.
Figure 14: Examples of applications using different voltage ranges
GaN-on-Si typical HEMT structures (Figure 15) consist of several epi-layers. These layers include
materials with a wider bandgap and a lower bandgap, and an AlGaN/GaN heterostructure. At the
interface between AlGaN and GaN, a two dimensional electron gas (2DEG) is created with an
electron channel accumulation without extrinsic doping. The 2DEG formation results from both
the spontaneous and piezoelectric effects [48], [54], [68]. The thickness and Al content of the
AlGaN barrier layer defines the resulting polarization. It can be pointed out that a high electron
mobility above 2000 cm2/Vs can be combined with a high carrier density within the 2DEG, thus
resulting in excellent electrical performance.
Figure 15: Cross section of a typical AlGaN/GaN HEMT structure
One of the main issues for GaN-based heteroepitaxy is the lattice mismatch and difference in
thermal expansion coefficients with the substrate. This generally leads to a high dislocation
density, which may be a source of leakage current under high electric field and subsequent device
degradation. Residual stress may be created, inducing eventually cracks. The most common
materials used as substrates are silicon, sapphire, silicon carbide and more recently bulk GaN. In
all cases, the epitaxial layers have rather high dislocation density (103 - 1010 cm-2 [69], [70]). Some
properties of the substrates typically used for GaN-based epitaxy are shown in Table 5. For power
applications, the Si substrate is preferred because of its low cost and availability in large diameter
(up to 12 inches), despite a 17% lattice mismatch and a strong difference in thermal expansion
coefficient between GaN and Si, which makes the growth challenging. Although sapphire
substrates combine a high resistivity and low cost, the low thermal conductivity leads to a
significant self-heating, which is not suitable for power applications. Finally, the high cost of
silicon carbide is prohibitive for large volume applications despite its outstanding properties.
Emerging GaN substrates, perfectly lattice-matched, are of interest for vertical GaN architectures,
see also Section 7.
In order to grow crack-free and high quality GaN films by reducing the defect density, especially
when using silicon substrates, the tensile stress during the growth and cooling process needs to be
limited. An AlN nucleation layer (NL) is typically used as an initiating layer for GaN growth. By
using a AlN NL, the melt-back etching of Ga into Si can be avoided [71]. Besides, the AlN NL
provides a GaN layer with a compressive strain due to the 2.5% lattice mismatch between AlN and
GaN. This is necessary for compensating the tensile stress generated during the cooling process.
Figure 16 shows a TEM image of the interface between the Si substrate and the AlN nucleation
layer. Dislocations are reduced but still present across the buffer layers.
Whatever the choice of substrate, the buffer layers are critical. The high 2DEG sheet charge density
in GaN-based HEMTs enables significant drain current densities. Inadequate carrier confinement
within the channel leads to soft pinch-off characteristics and high sub-threshold leakage. The
presence of a high defect or impurity density in the buffer produces high leakage currents and poor
device reliability. Consequently, on top of a high quality AlN NL, a proper buffer configuration
and material quality is mandatory.
A well-known approach is the use of a step graded AlGaN buffer, as shown in Figure 17. Several
micrometers thick AlxGa1−xN buffer layers with various Al-content enable to further mitigate
lattice and thermal-mismatch detrimental effects. The interest of this approach is also to improve
the 2DEG electron confinement under high electric field, by limiting the carrier injection into the
buffer layers, i.e. to the so-called punch-through effect. Another practical route to significantly
increase the buffer layers’ resistivity is the introduction of intentional compensating centers. Iron
or carbon doping can be used to produce highly resistive buffers (carbon being the choice for GaN-
on-Si GaN devices). Finally, super-lattice (SL) buffers consisting in many periods of thin
AlN/GaN pairs have been proved to be one of the most effective techniques both to control the
stress and enhance the blocking voltage [74]. With thick SL buffers, a high crystal quality and
smooth surface of top-GaN layer can be obtained resulting in superior performance as described
further in this paper.
Figure 17: Schematic cross section and TEM image of AlGaN layers with various Al content
between the nucleation layer and the GaN layer. Reprinted from "AlGaN/GaN HFET grown on 6-
inch     diameter    Si(111)    substrates   by    MOCVD",        S.  M.   Cho    et   al.,
https://doi.org/10.7567/SSDM.2011.AL-7-3, licensed under CC BY [75].
The epitaxial heterostructure is completed with a cap layer, which is generally composed of GaN
or SiN to reduce the oxidation of the underlying AlGaN film. It is important to note that surface
states [55] on top of the AlGaN constitute the origin of the 2DEG, as discussed in Section 5.
However, surface defects may also be detrimental to the device performance. Under operating
conditions, trapping at surface defects can create a virtual gate between gate and drain terminals,
depleting unintentionally the 2DEG and thus severely degrading the device performances and/or
reliability. Negatively charged surface states may compensate the donor atoms, thus depleting the
channel between the gate and the drain. A surface passivation layer, typically SiN, allows to
mitigate the 2DEG depletion as can be seen in Figure 18.
Figure 18: Schematic cross section showing surface state depletion effects within an AlGaN/GaN
HEMT
Lateral heterostructure devices are inherently normally-on, i.e. they conduct current when no gate
voltage is applied. This raises safety concerns because in case of a malfunctioning gate driver, the
GaN transistor is not automatically switched off and an uncontrolled current flow can damage the
entire system. Furthermore, normally-on transistors make circuit designs more complex because a
negative-voltage supply is required. Thus, a substantial research effort was focused on creating
normally-off devices in recent years.
In a conventional AlGaN/GaN HEMT, the threshold voltage VTH depends on several parameters
related to the gate metal and the heterojunction properties, as can be seen from the following
equation [76]:
                                                   𝜎(𝑥)                   𝑞𝑁𝐷
                 𝑉𝑇𝐻 (𝑥) = ɸ𝐵 (𝑥) − ∆𝐸𝐶 (𝑥) −                   𝑡−                  (t)²
                                                𝜀0 𝜀𝐴𝑙𝐺𝑎𝑁 (𝑥)        2𝜀0 𝜀𝐴𝑙𝐺𝑎𝑁 (𝑥)
𝑥 represents the Al content in the barrier layer; ɸ𝐵 (𝑥) is the Schottky barrier height between the
gate metal and the AlGaN barrier layer; ∆𝐸𝐶 (𝑥) is the conduction band discontinuity at the
AlGaN/GaN interface; 𝜎(𝑥) is the polarization charge at the AlGaN/GaN interface; 𝜀0 is the
permittivity vacuum; 𝜀𝐴𝑙𝐺𝑎𝑁 (𝑥) is the permittivity of the AlGaN layer; 𝑡 is the AlGaN thickness;
𝑞 is the electric charge; 𝑁𝐷 is the doping.
Thicker AlGaN barriers and higher polarization differences between AlGaN and GaN lead to more
negative VTH as they increase ns at zero bias. From this equation, it is clear that several degrees of
freedom exist to tune the VTH, such as changing the Schottky barrier height or the 2DEG carrier
density related to the AlGaN barrier layer, which is dependent on the Al content and its thickness.
Different topologies have been proposed in order to achieve normally-off GaN HEMTs: a cascode
configuration [77]–[79] combining a silicon normally-off MOSFET and a normally-on GaN
HEMT, the use of a HEMT with fluorine implantation under the gate [80]–[82], a gate recessed
MISHEMT (metal-insulator-semiconductor) with partial [83], [84] or complete [85] AlGaN
barrier removal, and a p-GaN-gated [24], [86], [87] HEMT.
Therefore, the cascode configuration enables to take advantage of the positive threshold voltage
of the MOSFET as well as the low on-resistance of the 2DEG, together with the high breakdown
field of the GaN HEMT in off-state conditions. However, it can be noticed that this approach limits
the high temperature operation by the presence of the Si device. In addition, the packaging
complexity and size are increased and parasitic inductances are introduced, and this may have an
impact on the switching performance of the circuit.
The choice of the dielectric is extremely important, as it will directly impact the channel mobility
within the 2DEG [88] and the stability of the threshold voltage [89]. Also, the dielectric quality
and surface roughness of the etched area are critical parameters and the interface charge density
needs to be well controlled. Several mechanisms, involving the surface states and related trapping
have been proposed to explain the possible origin of the device degradation phenomena (see
Sections 8.2 and 8.3).
Also, a high Mg concentration in the p-GaN layer is required, which should be balanced with the
deterioration of the crystal quality for too high Mg doping concentration. The p-GaN layer has
shown a wide process window in terms of thickness and doping, which eases process control
requirements. However, the low selectivity of the etching process between the p-GaN and the
barrier layer needs to be carefully optimized in order to achieve a stable high threshold voltage
[92].
Figure 24: Simulated conduction band diagrams of a p-GaN/AlGaN/GaN heterostructure for two
different Al content (12% and 26%) into the AlGaN barrier layer (left) and two different AlGaN
barrier thickness (right). Reprinted from Materials Science in Semiconductor Processing, Volume
78, G. Greco, F. Iucolano, F. Roccaforte, "Review of technology for normally-off HEMTs with p-
GaN gate", Pages 96-106, Copyright 2018, with permission from Elsevier [92].
Figure 25: Device operation modes as a function of the Al content and the thickness of the barrier.
Reprinted from Materials Science in Semiconductor Processing, Volume 78, G. Greco, F.
Iucolano, F. Roccaforte, "Review of technology for normally-off HEMTs with p-GaN gate", Pages
96-106, Copyright 2018, with permission from Elsevier [92].
Another parameter not to be neglected is the choice of the gate metal. The Schottky barrier height
between the metal and the p-GaN layer directly impacts the threshold voltage and follows the
relation:
                                            𝐸𝑔
                                     ɸ𝐵 =      − (ɸ𝑚 − χ𝑠 )
                                            q
where ɸ𝐵 is the Schottky barrier height between the gate metal and the AlGaN barrier layer; 𝐸𝑔 is
the bandgap of the semiconductor; ɸ𝑚 is the work function of the metal; and χ𝑠 is the electron
affinity of the semiconductor.
Based on this equation, metals with a lower work function provide a higher barrier height [97].
6.2.5 Tri-gate
Compared to conventional planar AlGaN/GaN HEMTs, tri-gate transistors feature fins patterned
in the gate region which are conformably covered by a 3D gate electrode (Fig. 26(a-b)). While
such architecture was first introduced for Si ultra-scale MOSFET in 2002 [98] to address the short
channel effects, it was soon adopted for power devices. The first tri-gate GaN demonstration
appeared in 2008 [99], followed by several other works which showed the several benefits of such
architecture, such as the improved current stability, the lower subthreshold slope, and the reduced
off-state leakage [100]–[103].
Figure 26 Schematics of (a) the tri-gate MOSHEMTs and (b) its tri-gate region (c) Dependence of
VTH on wfin. The inset in (c) illustrates the effect of sidewall depletion in distributing the 2DEG
across a fin [104]
The most interesting feature of the tri-gate is however its ability to modulate the device threshold
voltage (VTH) by simply tuning the fin width (wfin) (Fig. 26 (c)). Such effect has been attributed to
the partial AlGaN strain relaxation when patterned into fins and by the fin sidewalls depletion due
the side gate electrode [104]–[106].
The dependence of the device VTH on the fin width can be used to achieve normally-off operation
by simply designing the proper wfin by lithography, without the need for any critical etching as for
gate recess or p-GaN gates. However, early demonstrations based on this approach typically
presented still negative VTH (at 1μA/mm) and degraded on-resistance due to the very small fin
width required to achieve normally-off operation [100], [107]–[109]. Recently, a tri-gate device
with 20 nm-wide fins in combination with a large work-function gate metal stack (Pt/Au)
successfully showed full normally-off operation with a VTH of 0.64 V (at 1μA/mm) and a
competitive RON of 7.4 Ω·mm [110]. To achieve such performance, however, very high-resolution
lithographic processes are required both to define the fin width (which sets VTH) and to reduce the
gap in-between fins (which degrades RON). Since, at the moment, the lithographic resolution in
power devices foundries is still in the order of several hundreds of nm, it is interesting to find
solutions to increase the minimum required wfin.
A promising approach consists of combining the tri-gate architecture with conventional methods
to achieve normally-off operation such as recessed gate and p-GaN cap. This allows to achieve
large positive VTH values while keeping the benefits of the tri-gate architecture and relaxing the
lithographic requirements. The first demonstration of such an approach appeared in 2012 [111]
with the integration of the tri-gate architecture with AlGaN barrier recess, followed by a more
recent work [112] in 2019 which showed promising VTH of 1.4 V (at 1μA/mm) and RON of 7.3
Ω·mm. Besides, the tri-gate architecture can also be integrated with the pGaN cap approach,
allowing to further increase VTH with respect to the planar case and relax the trade-off between VTH
and the heterostructure sheet resistance [113]. Finally, the tri-gate architecture can also be
combined with promising p-type oxides, such as NiO, which can be conformably deposited around
the fin and help to further shift the device VTH without the need of complex pGaN regrowth [114].
Converters based on MOSFET have the ability to survive a limited exposure to voltages above the
device rating, according to the specified avalanche energy rating. However, lateral GaN HFETs
do not have the potential for avalanche breakdown [122], because they do not rely on a p-n junction
for blocking voltage and may experience catastrophic dielectric breakdown when exposed to
sufficient overvoltage [123]. This breakdown is destructive and non-recoverable.
The main sources of leakage current and related breakdown voltage (VBR) for an AlGaN/GaN
HEMT power transistor on Si substrate are the following (see Figure 27):
Different approaches have been developed in order to mitigate these leakage paths and associated
premature breakdown while avoiding trapping effects.
- The use of high-quality dielectric for surface passivation reduces leakage at the surface and at
the interface with the barrier [124].
- Proper doping compensation into the buffer layers, generally carbon or iron doping [125]–[127]
enables to significantly enhance the buffer resistivity and consequently avoid the carrier injection.
Furthermore, back barrier [128], [129] based on graded AlGaN or AlN material have been
developed in order to reduce this phenomenon, increasing the blocking voltage. In both cases, the
thickness of the GaN channel region must be carefully optimized with the aim of achieving a high
electron density in the 2DEG, a good electron confinement in the channel, and low trapping effects,
especially in the case of doping compensation [130], [131].
Figure 27: Schematic representation of the main sources of leakage current for AlGaN/GaN
transistors on Si
It can be pointed out that the AlGaN/GaN transistor breakdown voltage scales linearly for small
gate-drain distances (typically below 15 µm), while larger gate-drain distances result in a
saturation of VBR due to the conduction into the substrate triggered by the vertical electric field.
Extensive research is carried out to overcome the breakdown mechanisms and further push the
limits of GaN-on-Si HEMTs.
In addition, the crystal quality is an important factor. Considering the high material defect density
in GaN/Si epilayers, recent studies [132], [133] showed that the presence of defects may impact
on the breakdown voltage.
Figure 28 Schematics, equivalent circuits and distributions of potential (Φ) and electric field (E)
in lateral GaN transistors in OFF state with (a) no FPs, (b) a single FP, (c) two FPs and (d) a slant
FP [102]
Figure 29 Cross-sectional SEM images of (a) a single FP [147], (b) multiple FPs [148] and (c) a
slant FP [149] in GaN HEMTs
6.4.2 Buffer optimization: Super-lattice buffer
Unintentionally doped GaN buffer layers deliver insufficient resistivity for high voltage operation,
due to the residual n-type conductivity of GaN, which can induce parasitic leakage paths, thus
increasing the off-state leakage current. As previously mentioned, high resistivity can be achieved
by doping with deep acceptor impurities (such as C atoms), to compensate the background donors.
However, this approach can generate severe current collapse, if the buffer is not carefully
optimized [150]–[152].
To further improve the carrier confinement while suppressing undesirable trapping effects, the
doping compensation can be combined with the use of an AlGaN back barrier [74] or super-lattices
[153], [154] consisting in AlN/GaN pairs (see Figure 30). By alternating thin layers of high
crystalline quality wide-bandgap semiconductors (e.g. AlGaN, AlN or GaN), the accumulation of
internal stress can be minimized, thus creating a highly insulating buffer with low buffer trapping
effects.
Figure 30: Schematic cross section of a GaN HEMT using a super-lattice-based buffer and TEM
image of an AlN/GaN superlattice
                                                                            1400
                                                                                       REF
                                                                            1300       SL
 Figure 31: Vertical breakdown of a step-graded GaN Carbon doped buffer (REF) and a SL
 structure (black) at room temperature. Reprinted from "High Breakdown Voltage and Low
 Buffer Trapping in Superlattice GaN-on-Silicon Heterostructures for High Voltage
 Applications", A. Tajalli et al., https://doi.org/10.3390/ma13194271, licensed under CC BY
 4.0 [74].
A. Tajalli et al showed that the insertion of super-lattices (SL) into the buffer layers allows pushing
the vertical breakdown voltage above 1200 V without generating additional trapping effects as
compared to a more standard optimized step-graded AlGaN-based epi-structure using a similar
total buffer thickness (see Figure 31). dc characterization of fabricated transistors by means of
back-gating transient measurements reflect the much lower trapping effects and the benefit of the
SL (Figure 32).
Figure 33: Schematic cross-section of AlGaN/GaN MISHEMT including the front side process,
the LSR technique, a thick PVD AlN and metal backside deposition. Reprinted from "GaN-on-
silicon high-electron-mobility transistor technology with ultra-low leakage up to 3000 V using
local substrate removal and AlN ultra-wide bandgap", Ezgi Dogmus et al., Appl. Phys. Express,
https://doi.org/10.7567/APEX.11.034102, licensed under CC BY 4.0 [25].
E. Dogmus et al used the following device processing, which consists in ohmic contacts formed
directly on top of the AlGaN barrier by rapid thermal annealing. After device isolation, a metal-
insulator-semiconductor gate structure was employed by depositing Ni/Au metal stack on top of
the in-situ SiN cap layer. Once the front-side processing was completed, the Si substrate is locally
etched up to the AlN nucleation layer around the entire device (see Figure 33). Devices with and
without LSR have been fabricated on the same samples, eliminating any processing or epi variation
during the device characterization. Electrical characterization showed a slight decrease of the
maximum current density after LSR as can be seen in Figure 34 due to self-heating effects [33].
Further improvement of the heat dissipation would be required to avoid the decrease of the current
density. On the other hand, a drastic enhancement of the blocking voltage is achieved by locally
replacing the substrate with a wider bandgap material (Figure 35).
Figure 34: (a) Transfer and (b) output characteristics of GaN-based MISHEMTs with and without
LSR/backside AlN and Cu. Reprinted from "GaN-on-silicon high-electron-mobility transistor
technology with ultra-low leakage up to 3000 V using local substrate removal and AlN ultra-wide
bandgap", Ezgi Dogmus et al., Appl. Phys. Express, https://doi.org/10.7567/APEX.11.034102,
licensed under CC BY 4.0 [25].
Figure 35: (left) Evolution of LGD-dependent device VBR and specific on-resistance (inset) of
AlGaN/GaN HEMTs with and without LSR by defining the blocking voltage at ID =1µA/mm, and
(right) off-state leakage current characteristics of AlGaN/GaN MISHEMTs with and without LSR.
Reprinted from "GaN-on-silicon high-electron-mobility transistor technology with ultra-low
leakage up to 3000 V using local substrate removal and AlN ultra-wide bandgap", Ezgi Dogmus
et al., Appl. Phys. Express, https://doi.org/10.7567/APEX.11.034102, licensed under CC BY 4.0
[25].
Furthermore, the effects of Si removal were investigated by Raman thermometry [156], [157],
which revealed a worsening of the thermal performance. A significant improvement of the thermal
dissipation is obtained after the AlN and copper deposition.
Abid et al. used an AlN barrier on top of an Al50Ga50N channel grown on an AlN/sapphire template
(Figure 36). The heterostructure provides a high carrier concentration close to 1.9×1013 cm-2 with
a rather limited electron mobility of 145 cm²/V.s. Low leakage current is obtained without the use
of any field plates, confirming that tunneling mechanisms are not present in Al-rich transistors. In
general, Al-rich transistors are less prone to gate leakage than AlGaN/GaN HEMTs. Despite the
rather high defect density, a blocking voltage above of 4000 V with an off-state leakage current
below 0.1 µA/mm is achieved for AlGaN-based channel HEMTs (Figure 37). It can be noticed
that low gate-drain distance of 5 µm yields a breakdown field of 3.5 MV/cm, which is well-beyond
that of SiC and GaN devices. Furthermore, these transistors show a very stable behavior as a
function of temperature, with no threshold voltage variation and low off-state leakage increase up
to 200°C.
                1
                                                                                                 Regrowth
             100m   GD5                                                          100             Partially etched barrier
                    VDS = 4V
             10m                                                                     GD5
                                                                                  80 VGS [-16V; 0V]
              1m
                                                                                           by step of 2V
                                                                     ID(mA/mm)
 ID (A/mm)
             100µ                                                                 60
              10µ
                                      Partially etched barrier                    40
               1µ
                                      Regrowth
             100n
                                                                                  20
              10n
               1n                                                                  0
                 -20 -18 -16 -14 -12 -10   -8   -6   -4   -2     0                     0       2      4    6     8     10   12   14   16   18   20
                                   VGS (V)                                                                           VDS (V)
Figure 38: Transfer characteristics (left) and output characteristics (right) of AlN/AlGaN/AlN
HEMTs using regrown ohmic contacts and partially etched barrier
One of the major challenges limiting the research progress of Al-rich AlGaN transistors is the
optimization of ohmic contacts. High resistance of the source and drain electrodes, and the possible
Schottky-like behavior, lead to a reduced current density. Low resistance ohmic contacts are
fundamental performance enablers in wide bandgap HEMTs. Several approaches [162]–[165] are
under development to mitigate this issue, namely based on tuning the heterostructure and/or the
subsequent annealing. For instance, regrown ohmic contacts using a SiO2 mask can be applied.
The reduction of the contact resistances has been verified at the transistor level. Output
characteristics of transistors with regrown non-alloyed contacts showed a significant current
density increase above 100 mA/mm as compared to identical devices with partially etched barrier
and annealed contacts (see Figure 38). This is clearly resulting from the drastic drop of the contact
resistances. In spite of these obstacles, preliminary results show that contact resistivity will
improve over time and that advanced approaches such as the etch and regrowth processes will
ensure successful achievement at even higher Al-composition.
As the AlGaN/GaN power technology is reaching its maturity and increasing its market share, the
research on this topic is likely to follow two parallel paths. On one hand, there will be a growing
interest on the aspects directly related to the realization of a successful commercial product such
as trapping, reliability, stability, packaging, and circuit operation. On the other hand, researchers
will continue to come up with novel device structures and designs to improve performance and
take full advance of the GaN material properties for future generations of high efficiency devices.
Under this point of view, a few recent promising directions are presented below which are rapidly
growing and offer a significant advance in device performance.
                      B
                                                                                        wtrench wfin
                                                             (b)
                                                                   E-mode Slanted      D-mode Planar
                                                                    Gate Field plate Termination region
      Buffer layers
      Substrate
                                                            (c)                                  250 nm
 Figure 39: Three-dimensional schematics of the multi-channel power device, featuring multiple
 parallel channels, controlled 3-dimensionally by a tri-gate electrode. (b) FIB cross-section and
 schematics of the multi-channel nanowires covered by the tri-gate structure along the AB line
 in figure (a). The scale bar is 100 nm. (c) Top SEM image of the tri-gate area which includes,
 starting from the source side, an e-mode region achieved by 15 nm-wide nanowires, and a
 slanted region terminated on 100 nm-wide d-mode nanowires for optimal electric field
 management. Reprinted by permission from: Springer Nature, Nature electronics, "Multi-
 channel nanowire devices for efficient power conversion", L. Nela et al., Copyright 2021 [166].
Despite the recent progress, the performance of AlGaN/GaN devices is still far from the theoretical
limit predicted for the GaN materials [167]. A direct way to improve the device’s performance is
represented by increasing its carrier concentration, which directly leads to a reduced on-resistance.
Yet, achieving a large ns leads to major challenges for the heterostructure and device design. First,
a large ns severely impacts the mobility (μ) due to the increased electron-to-electron scattering,
limiting the reduction of the heterostructure Rsh. Secondly, a large ns leads to a difficult control of
the channel, which results in negative VTH and degrades the device voltage blocking capability.
                                                                                                                               53 A/mm
                     IME '16 PKU '18                                                  1 mA/mm                                                  SINANO '16        Samsung '12, 1 mA/mm
              8                                   CAS '19                            MIT '12
                                                                                                      NUS '12, 0.1 mA/mm                        1 A/mm
                                                                                                                                                                                EPFL '17
                  SYSU '17                                                         0.6 A/mm                                                       Panasonic '13
                                                                                                                                                                                45 A/mm                 it
                                                                                                                                                      1 A/mm                                         Lim
                                                                                                              NIT '13, 0.57 mA/mm                                                                iC
                                 EPFL '19       EPFL '19                                                                                               MIT '12,1 mA/mm                         -S
                                                                                                        NTU '11
                                                                                                      0.1 mA/mm
                                                                                                                           EPFL '17
                                                                                                                                                             EPFL '17
                                                                                                                                                                                            4H
                                                                                                                          0.3 A/mm
              6     HKUST '15                                                                                                                                1 A/mm
                                                                                                                             UESTC '15
                                                                               1   HKUST '13                                 1 A/mm            IQE '15
                                                                                                                                               1 mA/mm
                                                                                    1 A/mm
                                                                                                    EPFL '17, 0.3 A/mm
              4 Multi-channel                                                                                                                             Multi-channel @ 1 A/mm:         it
                                                                                                 Toshiba '13, 10 A/mm
                                                                                                                                                            E-mode                     Lim
                                       >2x lower RON                                                                                                                                 N
                                                                                        IQE '15, 1 mA/mm                                                                           Ga
                                                                                                                                                            D-mode
              0.0       0.5         1.0       1.5          2.0                                                                      1                                              2                          3
    (a)                         VTH (V)                                (b)                                                                VBR (kV)
 Figure 40 RON vs VTH benchmark for the multi-channel device against state-of-the-art power
 devices. VTH has been defined at 1 μA/mm. (c) RON,SP vs VBR benchmark for normally-off and
 D-mode Multi-channel devices with respect to state-of-the-art GaN-on-Si (MOS)HEMTs.
 Reprinted by permission from: Springer Nature, Nature electronics, "Multi-channel nanowire
 devices for efficient power conversion", L. Nela et al., Copyright 2021 [166].
While multi-channel devices have first been proposed for RF applications [172]–[175], there has
been a growing interest in their use in power electronics applications. However, power devices
present very specific requirements such as normally-off operation, large blocking voltage
capabilities and good stability during switching operation, which need to be separately addressed
and solved. The first power multi-channel power HEMT was reported in 2018 [176], followed by
the demonstration of a high-voltage multi-channel SBD [177]. While these early works showed
the concept of multi-channel power devices, their performance improvement was still quite limited
due to the relatively high sheet resistance (~ 240 Ω/sq) of the multi-channel heterostructure
employed. More recent works [166], [178], [179], however, showed the full potential of the multi-
channel technology for power devices. By employing a highly conducting multi-channel
heterostructure (Rsh of 83 Ω/sq) in combination with a carefully designed slanted tri-gate structure,
multi-channel power devices showing normally-off operation with VTH of 0.85 V (at 1 μA/mm),
on-resistance of 3.2 Ω·mm and breakdown voltage of 1300 V (at 1 μA/mm) were demonstrated
[166]. Such performance considerably surpasses the state-of-the-art of conventional single-
channel devices and opens new perspectives for GaN power devices (Figure 40). Besides, multi-
channel devices passivated by low-pressure chemical vapor deposition (LPCVD) Si3N4 presented
reduced current collapse up to high voltage stress and excellent VTH stability both during switching
and high-temperature operation, showing the potential of such technology [180]. Further research
on this topic will likely concentrate on the optimization of the multi-channel heterostructure and
on additional methods to achieve large positive VTH.
(a) (b)
 Figure 41 Schematic of the cross-section of a conventional power MOSFET (a) and a super
 junction device (b) with the corresponding electric field distribution under off-state conditions.
 Reprinted from: H. Kawai et al., "Low cost high voltage GaN polarization superjunction field
 effect transistors", Wiley & Sons, Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA,
 Weinheim [181].
The realization of conventional vertical SJ (Figure 41) in GaN is however not straightforward due
to difficult technological challenges, among which the inefficient Mg-based p-doping of GaN is
one of the most relevant. On the one side, the reduced Mg activation ratio results in relatively low
doping concentrations, which, combined with the difficult control of the exact doping level, make
charge-matching extremely challenging. On the other side, the absence of efficient implantation
doping and high-quality p-GaN regrowth, hinder the realization of the typical vertical SJ pillars.
These challenges make the demonstration of vertical GaN SJ devices still out of reach and, until
more efficient p-GaN doping is achieved, it is unlikely that this technology could progress
significantly.
 Figure 42 Schematic illustration of a PSJ structure. Reprinted from: H. Kawai et al., "Low cost
 high voltage GaN polarization superjunction field effect transistors", Wiley & Sons, Copyright
 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim [181].
Yet, alternative solutions to realize SJ devices have been proposed in AlGaN/GaN lateral
architectures. In these structures, it is possible to obtain a two-dimensional electron and hole gas
(2DEG and 2DHG) of equal concentration thanks to the presence of matching polarization charges.
Such devices are typically referred to as polarization super junctions (PSJs) (Figure 42) and can
result in similar behavior to conventional, doping-based SJs, thus yielding much-improved off-
state performance. While these devices were first proposed in 2008 [187]–[189], technical
challenges in achieving a good match between the 2DEG and 2DHG concentrations have slowed
the development of this technology. After some years, the research on this subject has regained a
strong interest and several works have recently appeared on this topic [181], [190]–[192],
providing further insight on the devices working principle and showing its potential. Fast
development of this field in the next years is thus expected, which could result in a new generation
of GaN super-junction devices.
7 Vertical GaN device structures
7.1   Why Vertical GaN?
The most common GaN-based power devices available presently are GaN HEMTs, which have
been discussed in detail in the previous sections. GaN HEMTs rated for 650 V/900 V breakdown
voltage (BV) and maximum output DC current (IDS) as high as 150 A are available commercially
[193]–[195] for a broad spectrum of applications, such as on-board battery chargers, high-
efficiency and high-density power converters, solar panel inverters, among many others.
However, the lateral topology of HEMTs presents some pertinent limitations related to reliability
and breakdown voltage scaling, which hinder the usability of these devices for high voltage
applications requiring breakdown voltages above 700 V, for example in electric and hybrid electric
vehicles (EVs and HEVs), photovoltaic (PV) inverters, wind turbines, traction systems for trains,
to name a few.
Most of the limitations associated with GaN HEMTs arise from the lateral electron flow between
the source and drain terminals, very close to the device surface. The density of electrons in the
channel is sensitive to the presence of surface traps, that may degrade the electrical performance
[196]–[198] of the transistors. This leads to issues like current collapse and dynamic degradation
of the on resistance, which are more severe with increasing the BV rating of the device (see
Sections 8.2 and 9). In addition, the lateral nature of the transport results in a very inhomogeneous
distribution of the electric field in the device, peaking in specific regions (e.g. the edge of the gate,
or of the field plate, on the drain side). This may enhance electron trapping at surface states, and
may lead to premature breakdown of the semiconductor and of dielectrics. This ultimately
degrades the forward and reverse performance of the device [198], [199], and limits its full voltage
blocking potential. For traditional GaN HEMTs, the BV is dictated by the gate to drain spacing
(provided this value is smaller than the drain-to-substrate BV) and thus larger breakdown voltages
require larger device sizes, which also increases the device cost. Another major concern is that
GaN HEMTs are in general normally-on devices: this is not desirable for power electronics
applications from a safety perspective and for simplicity of the gate drivers, which are currently
designed for normally-off devices. As discussed in Section 6.2, Several methods to achieve
normally-OFF operation have been developed including cascode [200]–[202] configuration (based
on GaN HEMT combined with a Si MOSFET), fluorine ion implantation [80], [94], [203] under
the gate region, recessing the barrier layer [83] in the gate region, by applying tri-gate structures
[101], [102], [110], [111], [178] to the gate region, among others. However, in many cases and
even for commercial devices, the threshold voltage (Vth) that can be achieved is only around 1-2
V, which may not be ideal for fail-safe operation. Finally, GaN HEMTs do not present avalanche
capability, which can prevent device failure under short term over-voltage conditions. Thus, for a
GaN HEMT to qualify for a certain BV rating, the device has to be overdesigned to sustain a much
higher BV, which increases the device size and cost.
Vertical GaN power devices are different from their lateral counterparts as the current flows
vertically, i.e., parallel to the growth direction of the epitaxial GaN layers. The vast majority of
the Si and SiC power devices available commercially are based on this design philosophy and are
capable of delivering high ON-state currents (> 7000 A) and high BVs (> 8000 V) [204]. For
GaN, the vertical topology offers distinct advantages over lateral power HEMTs. The BV can be
increased by increasing the thickness of the voltage blocking layer, generally formed by an un-
intentionally or low doped GaN layer (also referred to as the drift layer or i-GaN layer),
independent of the size of the device. The RON in PiN diodes increases only slightly [205] with
increasing the thickness of the drift layer as a result of extrinsic phenomena like conductivity
modulation [206], [207]. Vertical devices are also not affected by surface traps as in GaN HEMTs,
and the electric field peaks well inside the GaN layers, away from the surface, thus improving the
device breakdown voltage and reliability. In addition, vertical MOSFETs can provide high positive
Vth of 5-15 V, which is well suited for higher power applications like in automobiles. Another
major advantage is the existence of avalanche breakdown [208]–[210] for GaN vertical power
devices. This greatly improves the reliability, and eliminates the need to overdesign the device.
The ideal solution for obtaining high quality GaN epitaxial layers is by homo-epitaxy i.e., GaN
layers grown on bulk GaN substrates. These substrates are mainly produced by hydride vapor-
phase epitaxy (HVPE), although several other methods like Na-flux or ammonothermal growth
are currently being investigated [211]–[216]. The major advantages of growing GaN by
homoepitaxy are their low dislocation density of 104 -106 /cm2 and the inherently matched lattice
and coefficient of thermal expansion (CTE) to the grown GaN layers. As a result, thick GaN layers
suitable for achieving high-voltage (~ 5000 V) [217], [218] vertical power devices can be easily
grown on these substrates. The downside is that these substrates are very expensive, at about 50
$/cm2, and mostly available in small 2-inch wafer diameters [219]; strategies for larger size
substrates are currently under investigation [211], [220] []. This hinders the widespread
commercialization of devices grown on bulk GaN. The bulk GaN market is also highly
concentrated, as three companies based out of Japan hold about 85% stake in the bulk GaN market
[221]. Currently, these expensive substrates are being used only for special applications, like laser
diodes and high-brightness LEDs [221], for which low dislocation densities are essential. Hence,
in order to take advantage of the material benefits offered by GaN materials for power device
applications, further improvements in wafer size and reduction in cost are highly desirable.
However, over the past decade, improvements in wafer size have been relatively slow, which is a
critical aspect for reducing the production cost per device. A strategy to tackle this issue would be
by hetero-epitaxy i.e., GaN layers grown on foreign substrates like Si, SiC and sapphire, which are
cheaper than bulk GaN and are available up to 12-inch diameters [222], [223]. However, these
substrates are both lattice and CTE mismatched to GaN, as shown in Table 6. This results in a high
defect density in the GaN crystalline structure as a result of the stress built up during growth.
Growth of thick layers of GaN (> 7µm) on 6-inch Si substrates also results in significant wafer
bowing and cracking [219], [224], [225]. Thus, further improvements in the dislocation density
and investigation of stress-relaxation buffer layers for the growth of thick GaN layers on these
substrates are essential.
Defect density ~ 109 /cm2 ~ 5×108 /cm2 ~ 109 /cm2 ~ 104-106 /cm2
CTE mismatch % 54 25 34 0
The majority of the reported GaN vertical power devices are based on bulk GaN substrates as a
result of their low dislocation density, which provide a fair representation of the superior material
properties of GaN as compared to Si. Even though these bulk GaN substrates have a defect density
higher than Si or SiC, these devices have been shown to pass reverse leakage tests, high
temperature reverse bias (HTRB), high temperature operating life (HTOL), temperature humidity
bias (THB), temperature cycling (TC) and inductive avalanche ruggedness stress tests [227], which
is of paramount importance from the point of view of commercialization of these devices in the
future. This section aims to summarize the research development on vertical GaN devices,
providing an extensive review encompassing the fabrication and performance of various vertical
devices reported till today. The development of vertical devices on sapphire and bulk GaN will be
first presented followed by the recent development of GaN-on-Silicon vertical devices.
P-i-N junctions are ubiquitous structures that compose many electronic devices. In wide-band-gap
semiconductors however, the large turn-on voltage and high reverse recovery times of such diodes
due to the large band gap hinders their applications in efficient power electronics. Even though
Schottky diodes are preferable in power applications, due to their low turn-on voltage and reverse
recovery time during switching transients, PN junctions are also an integral part of a number of
modern vertical power devices, including IGBTs, junction barrier Schottky (JBS) diodes, merged
p-i-n Schottky diodes, junction termination extensions (JTEs), etc. and thus merits a
comprehensive study. Furthermore, the well-known and simple physics of p-n junctions is helpful
in elucidating various material parameters, such as critical electric field (Ec), doping density,
impact ionization coefficients, generation-recombination rates, mobility of electron and holes,
temperature related effects, etc. [228] These properties are of utmost importance for the design
and understanding of the power device.
                     (a)                             (b)
Figure 43 (a) Schematic cross sections of the GaN p-n junction diodes with SiNx passivation and
the FP structure (https://ieeexplore.ieee.org/abstract/document/6042353) [229]. (b) Schematic
cross sections of the GaN p-n junction diodes with the triple drift layers and the FP structure.
(https://ieeexplore.ieee.org/document/7273835) [218].
The first report on GaN power p-i-n diodes on sapphire substrates date back to 2000 [230], [231]
and the first GaN p-i-n diodes on bulk GaN substrates were reported in 2005 [232]. However,
high-voltage p-i-n diodes with BV > 1 kV were reported only in 2011 [233]. Rapid developments
in the growth and fabrication of p-i-n diodes ensued, mainly by startups like Avogy Inc. and
researchers from institutions like Cornell University, Hosei University, Toyoda Gosei, etc [228],
[234]–[237]. The schematic of a p-i-n diode with BV of 1100 V [229] as reported by Hosei
university is shown in Figure 43 (a). A mesa termination with SiO2 passivation and a field plate
structure were employed to improve the BV from ~ 450 V to 1100 V along with a small R on,sp
(given by RON × Active area of the device) of 0.4 mΩcm2, thus achieving an excellent Baliga’s
                            𝐵𝑉 2
figure of merit (BFOM = 𝑅           ) > 3.0 GW/cm2. Subsequently, a low-damage field plate process
                            𝑜𝑛,𝑠𝑝
involving the use of a bilayer spin-on-glass (SOG)/ sputtered SiO2 as the field plate dielectric along
with a drift layer thickness of 20 µm were developed to achieve the first demonstration of GaN p-
i-n diodes with very high BV of over 3 kV [235]. The SOG protected the p-GaN anode contact
area from damages related to the SiO2 sputtering process. A low Ron,sp of 0.9 mΩcm2 along with
this high BV resulted in record BFOM of 10 GW/cm2 in 2013. A p-i-n diode [218] with triple drift
layer to improve the distribution of the electric field was proposed in 2015 (Figure 43 (b)). The
drift layer forming the p-n junction was doped to low 1015 /cm3 to create a near-flat electric field
profile and thus reduce the electric field. Subsequent drift layers were moderately doped to reduce
the Ron,sp to 1.7 mΩcm2, while still presenting a high BV of 4.7 kV. In 2018, a novel p-i-n diode
with a guard-ring termination [217] was presented which resulted in lower leakage current and an
improvement in BV by 200 V to obtain a high blocking voltage of 5 kV.
(a) (b)
Avogy Inc. first reported on the avalanche capability in p-i-n diodes with BV of 2.6 kV and 3.7
kV [228], [234] (Figure 44). The device structure is as shown in Figure 45 (a). An ion-
implantation-based proprietary edge termination was employed for better redistribution of electric
field peaks to realize breakdown voltages approaching 85% of that in theoretical parallel-plane
junction breakdown, and also to achieve avalanche capability. The role of substrate orientation on
the reverse leakage current and reliability of the devices revealed that a slight miscut angle of
several tenths of a degree is very beneficial. This results in the elimination of hillocks on the
surface of the as-grown GaN layers and reveals a surface with a smooth morphology which is
essential for achieving reliable devices with low reverse leakage currents [228] (Figure 45 (b)).
These devices were qualified with HTRB, THB, HTOL and TCT tests, indicating that even with a
defect density of 104-106 /cm2, GaN vertical devices could be adopted for fast commercialization
[227]. Avogy p-i-n diodes also fared extremely well against Si fast diodes when used in power
converter topologies, like the hard-switched boost circuit with little or no ringing as well as no
reverse recovery loss as compared to the Si fast diodes [238], [239]. They also demonstrated large
area 16 mm2 PiN diodes with current capability of 400 A in pulsed operation [240]. These diodes
provided 100 A at 4.5 V forward bias in DC operation together with an excellent BV of 700 V.
                                   (a)
(b)
Figure 45 (a) Schematic of a vertical GaN p-n diode with ion-implanted edge termination
(https://ieeexplore.ieee.org/document/6919319) [228]. (b) Nomarski image of surface morphology
observed on devices grown on GaN substrates. The image on the left demonstrates hillocks formed
on the GaN surface in growth on low miscut angle substrates. Devices fabricated using substrate
B will consistently have lower reverse leakage currents compared to those on substrate A.
Reprinted from Microelectronics Reliability, Volume 55, Isik C. Kizilyalli et al., "Reliability
studies of vertical GaN devices based on bulk GaN substrates", Pages No. 1654-1661, Copyright
2015,                  with              permission                from                Elsevier.
(https://www.sciencedirect.com/science/article/pii/S0026271415301013) [227].
Cornell University demonstrated the growth of high-quality GaN layers by metal organic chemical
vapor deposition (MOCVD) resulting in a Shockley-Read-Hall (SRH) lifetime of 12 ns. As a
consequence, their p-i-n diodes exhibited ultra-low Ron,sp of 0.12 mΩcm2 coupled with a high BV
of 1.4 kV, thus resulting in a BFOM of 16.5 GW/cm2 [241]. GaN p-i-n diodes incorporating a
bevel termination and a long field plate were also presented in 2015 with an improvement in BV
by more than 3-fold to ~ 4 kV[242], as compared to a p-i-n diode with no termination. In order to
circumvent the issues faced during p-GaN growth by MOCVD, like the Mg memory effect and
the hydrogen passivation of Mg dopants in p-GaN, high BV GaN p-i-n diodes with molecular
beam epitaxy (MBE) grown p-GaN were also reported [243], [244]. This study also provides an
alternative strategy for p-GaN regrowth by MOCVD, which could result in impurity incorporation
at the growth interface and issues arising from non-planar growth, like high leakage currents [236],
[237], [245].
                (a)                                        (b)
Figure 46(a) SIMS profile of the regrown p-n junction showing high levels of Si and O atoms at
the regrowth interface. Reprinted from K. Fu et al., "Investigation of GaN-on-GaN vertical p-n
diode with regrown p-GaN by metalorganic chemical vapor deposition" , Applied Physics Letters
113, 233502 (2018), https://doi.org/10.1063/1.5052479, with the permission of AIP Publishing.
(https://aip.scitation.org/doi/full/10.1063/1.5052479) [236]. (b) Schematic cross-section of GaN-
on-GaN      p-n       diodes    with     hydrogen       plasma      based     edge      termination
(https://ieeexplore.ieee.org/document/8360159). [246]
Arizona State University reported on the beneficial effects of growing a thick buffer layer of about
1 µm on the bulk GaN substrate prior to subsequent growth [247]. They also investigated in details
the regrowth of p-GaN layers on etched GaN surfaces [236], [248]. Regrowth of p-GaN is an
important topic, especially since ion implantation schemes for the realization of JTE (junction
termination extension) structures similar to Si and SiC carbide vertical power devices are very
complicated in GaN and still not available. The study revealed that the regrowth process could
result in a slight increase in the edge dislocations and a high concentration of Si and O impurity
atoms at the growth interface (Figure 46(a)). In a subsequent work [248], the reason for this high
impurity atom concentration was shown to be from the defective etching process, and a low power
etching coupled with UV-ozone/acid treatment of the etched surface prior to the p-GaN regrowth
was presented as a remedy to this issue. A novel edge termination scheme was also introduced by
passivating the p-GaN around the anode region by hydrogen plasma [246], to obtain significant
gains in the BV (Figure 46 (b)).
                           (a)               (b)
Figure 47(a) Schematic of the heterostructure used for the epitaxial liftoff devices clearly showing
the i-InGaN release layer (https://ieeexplore.ieee.org/document/8457225). [249] (b) Schematic
cross-section of GaN-on-GaN p-n diodes with nitrogen implanted edge termination Reprinted from
J. Wang et al.,"High voltage, high current GaN-on-GaN p-n diodes with partially compensated
edge       termination",         Applied     Physics     Letters      113,      023502        (2018)
https://doi.org/10.1063/1.5035267,         with    the   permission      of     AIP      Publishing.
(https://aip.scitation.org/doi/full/10.1063/1.5035267). [250]
Besides these developments, several other groups demonstrated important advances in the growth,
termination and improvements to the on-state electrical performance. University of Notre Dame
and Sandia National Laboratory demonstrated the use of nitrogen implantation to form edge
termination for p-i-n diodes [250], [251] ( Figure 47(b)). Devices with ultra-low Ron,sp of 0.15
mΩcm2 and BV of 1.68 kV corresponding to Baliga’s figure of merit of 18.8 GW/cm 2 were
obtained. Another type of termination is a bevel edge termination, which was studied extensively
for GaN p-i-n diodes by researchers from Kyoto University [210]. The study set forth important
design instructions, particularly the angle of the bevel, the thickness and the doping of the GaN
layers, etc.
Bulk GaN substrates are typically around 350-400 µm thick and serve the purpose of providing a
lattice and CTE matched template for epitaxial growth of GaN layers. However, after the growth
of the low defective GaN layers, the bulk GaN substrate could be effectively removed and reused.
This concept was demonstrated by introducing a thin i-InGaN layer between the bulk GaN
substrate and a p-i-n heterostructure [249] (Figure 47(a)). This i-InGaN layer can be photo-
electrochemically etched resulting in an epitaxial liftoff from the bulk GaN substrate. The epitaxial
layers can then be bonded to a high thermal conductivity material and processed further. The bulk
GaN substrate can then be reused many times for growing new epitaxial layers by a similar process.
The p-i-n diodes fabricated by using this method demonstrated an excellent Ron,sp of 0.2-0.5
mΩcm2 and a BV of 1300 V similar to a control device fabricated without epitaxial lift off, thus
confirming no degradation in the performance as a result of this special fabrication process.
These results revealed the excellent progress made for GaN PiN diodes providing important
insights into the material properties like critical electric field, reliability, avalanche capability, etc.
Si and SiC devices normally employ junction termination extension structures formed by selective
p/n- type doping by ion implantation to achieve reliable devices. Due to the difficulties involved
in achieving selective doping in GaN, several other methods based on N2/H2 plasma treatment,
field plate deposition and ion implantation schemes to form resistive regions around the anode
were employed to obtain high BV devices.
Unlike PiN diodes, SBDs are unipolar devices, which find applications in power converters by
virtue of their low turn-on voltage as well as absence of reverse recovery charge. Due to the
absence of a p-type layer, and thus conductivity modulation, SBDs normally have a higher RON
compared to PiN diodes [252]. The reverse leakage current is also higher as the reverse voltage is
held by the depletion at the metal-semiconductor Schottky barrier. Vertical GaN SBDs on sapphire
substrates were first reported in 2000 with a BV as high as 550 V [253], [254]. In 2001, vertical
GaN SBDs on bulk GaN substrate were demonstrated by incorporating a Mg+-ion implanted p-
guard rings at the edge of the Schottky contacts. The device presented a Ron,sp of 3.01 mΩcm2 and
a BV of 700 V. The first high voltage SBDs on bulk GaN substrates, presenting BV > 1 kV were
only demonstrated in 2010 by Sumitomo electric industries [255]. The growth was optimized
leading to an excellent electron mobility of 930 cm2/Vs in the undoped GaN layer resulting in an
ultra-low Ron,sp of 0.71 mΩcm2. A field plate termination was employed to obtain an excellent BV
of 1100 V using just a 5 µm-thick undoped GaN as the voltage blocking layer, leading to a BFOM
of 1.7 GW/cm2 (Figure 48(a)). Large-area SBDs with 1.1 × 1.1 mm2 Schottky electrode area
provided a forward current of 6 A at 1.46 V and a Ron,sp of 0.84 mΩcm2 while still exhibiting a
high BV of 600 V. This revealed the potential for scaling up these devices for commercial
applications. The switching characteristics of these large area diodes compared against Si fast
recovery diodes (FRDs) and SiC SBDs, which revealed the smallest reverse recovery time, reverse
recovery charge and losses for GaN SBDs [256] (Figure 48(b)). Their study also confirmed stable
forward and reverse aging characteristics for 1000 hours. Further advances in the material quality
of the bulk GaN substrates and optimizations in the MOCVD growth of GaN epitaxial layers led
to improvement in the electrical characteristics of SBDs, with forward currents of 50 A at 2.05 V
while sustaining a high enough BV of 790 V, as reported by Toyoda Gosei [257]. The 3 × 3 mm2
SBDs included a mesa termination with a field plate to improve the BV, while an excellent electron
mobility of 1200 cm2/Vs for the undoped GaN layer provided a low differential ON-resistance of
25-29 mΩ.
(a) (b)
Figure       48:   (a)   Schematic     cross    section    of    the    vertical    GaN      SBD
(https://ieeexplore.ieee.org/document/6856038). (b) Reverse recovery characteristics of GaN
SBD, SiC SBD, and Si FRD at IF of 5A, a reverse voltage of 380V, and dI/dt=3.4kA/sec.
(https://ieeexplore.ieee.org/document/6856038). [256]
From 2015-2020, significant progress was made in improving the quality of the epitaxial GaN
layers and termination methods for SBDs. Cao et al., from HRL laboratories demonstrated an SBD
with graded AlGaN cap layer on top of the voltage blocking drift layer (i-GaN) [258] as shown in
Figure 49.
                      (a)                            (b)
Figure   49    (a)   Schematic     cross   section    of   the   control   vertical   GaN     SBD
(https://aip.scitation.org/doi/full/10.1063/1.4943946) and (b) the SBD with the graded AlGaN cap
layer (https://aip.scitation.org/doi/full/10.1063/1.4943946) [258]
This reduced the reverse leakage current by three orders of magnitude while the polarization field
in the graded AlGaN effectively shortened the depletion width leading to the formation of a
tunneling current at relatively lower bias, thus providing a low turn-on voltage to 0.67 V. These
AlGaN capped SBDs with a Schottky contact area of 0.8 × 0.8 mm 2 improved the BV by more
than 2-fold, to 700 V, as compared to a control SBD with no AlGaN cap layer. The effect of C
incorporation during the MOCVD growth was investigated on both the forward as well as the
reverse characteristics of SBDs [259]. By varying the growth pressure and V/III ratio, different C
concentrations from  3 × 1015 to 3 × 1019 /cm3 could be obtained. Their study revealed that lower
C incorporation is better for both forward and reverse performance, resulting in SBDs with small
turn-on voltage of 0.77 V and high BV of 800 V for a large Schottky contact area (0.8 × 0.8 mm2)
devices. Prof. Amano’s group at Nagoya University also studied the effect of C impurity
accumulation on the leakage current of GaN SBDs [260]. In their study, initial failure of SBDs at
low voltages were ascribed to leakage current path through polygonal pits created by C impurity
accumulation during the growth process. Their group also perfected the method of achieving low
impurity levels in m-plane GaN, approaching c-plane values by using a quartz-free flow channel
[261]. In 2019, they reported on the effect of drift later thickness on BV, along with the
demonstration of vertical GaN SBD with the highest reported BV of 2.4 kV for a drift layer
thickness of 30 µm [262]. A reduction in the effective donor concentration with increasing the
thickness of the drift layer was observed by SIMS, and is believed to have a positive effect on
achieving such a high BV. In a recent publication, their group presented their results on SBDs with
a drift layer compensated for the un-intentional n-type doping by introducing Mg dopants during
the growth [263]. The resulting SBDs provided more than 3x-higher BV as compared to a non-
compensated SBD, but the on-state current and Ron,sp suffered as a result of the high resistivity of
the drift layer. Arizona State University investigated ways to balance the interplay between Ron,sp
and BV with the introduction of double drift layers (DDL) for SBDs [264]. Basically, the SBDs
consisted of an unintentionally-doped (UID) drift layer on the top and a slightly doped drift layer
at the bottom. The UID drift layer at the top could suppress the peak electric field at the Schottky
metal – UID interface and thus improve the BV.
Several implantation-based termination methods were demonstrated in the recent years. Han et al.
demonstrated a planar nitridation based-termination, by subjecting the area around the Schottky
contact to a N2 plasma from a plasma-enhanced chemical vapor deposition system (PECVD) [265].
From ultraviolet photoelectron spectroscopy, it was inferred that the Fermi level at the GaN surface
which underwent the N2 plasma treatment, went down by 0.68 eV possibly by the passivation of
the Ga dangling bonds and thus, an enlarged energy barrier height and/or effective barrier thickness
is presented at the junction edge. This suppressed the thermionic field emission (TFE)/ tunneling
at this region, and the leakage current reduced as a result. The SBDs with this termination scheme
presented 4 orders of magnitude lower leakage current as compared to a control SBD with no
termination and improved the BV from 335 V to 995 V. These devices also provided excellent
switching behavior with current-collapse free operation and zero reverse recovery characteristics
[266]. Han et al., also demonstrated fluorine implanted edge termination schemes for GaN SBDs
based on the principle that the implanted fluorine ions act as fixed negative ions in GaN [267]
(Figure 50(a)). This results in spreading of the electric field and prevents the localized peaking at
the Schottky contact edge, thus improving the BV as shown in Figure 50(b). The implantation was
done at energy levels of 30, 60 and 100 keV followed by post implantation annealing at 450 °C in
N2 ambient for 10 mins. The BV was boosted from 260 V for the unterminated SBD to 800 V.
Further improvement in the BV to 1020 V was achieved by capping the drift layer with a thin 5
nm layer of graded AlGaN, similar to that reported by Cao et al. Wang et al., demonstrated an
identical edge termination scheme with boron implantation [268]. Similar to the fluorine implanted
device described before, the boron implanted SBD provided 5 orders of magnitude improvement
in the leakage current and improved the BV from 189 V for the unterminated SBD to 585 V.
       (a)                                     (b)
Figure 50 (a) Schematic cross section of the control vertical GaN SBD with fluorine implanted
termination (https://ieeexplore.ieee.org/abstract/document/8709951) and (b) show simulated
electric field distribution in the unterminated-SBD and FIT-SBD at −600 V, respectively
(https://ieeexplore.ieee.org/abstract/document/8709951). [267]
(a) (b)
Figure 51(a) Schematic device top view and cross section of the fabricated trench JBS diode
(https://ieeexplore.ieee.org/abstract/document/7859485) and (b) show simulated electric field
distribution at -200 V clearly showing the reduction in electric field due to RESURF action from
adjacent p-GaN later (https://ieeexplore.ieee.org/abstract/document/7859485). [269]
             (a)                                (b)
Figure 52: Schematic cross section of a JBS diode by (a) Mg ion implantation in n-GaN
https://ieeexplore.ieee.org/document/7959570) and (b) Si           ion   implantation in     p-GaN
(https://ieeexplore.ieee.org/document/7959570) [270].
Junction Barrier Schottky (JBS) diodes are another type of device that combines low turn-on
voltage from the Schottky contacts and low leakage current provided by p-i-n diodes. This is
achieved by having alternate undoped and p-type regions below the anode contact which can be
easily formed for Si and SiC by ion implantation methods. Since ion implantation of dopants to
GaN is very difficult, other methods to achieve JBS structures were pursued. Cornell University
created a trench JBS diode by selectively etching away portions of p-GaN layer form a p-i-n diode
followed by Schottky contact formation [269], as presented in Figure 51(a). The reduced surface
field (RESURF) effect at the Schottky surface as a result of the adjoining p-GaN layers were
elaborately studied using TCAD (Figure 51(b)). Hayashida et. al. from Mitsubishi Electric
Corporation demonstrated a merged p-i-n Schottky diode [271] based on a trench JBS diode
achieving a BV of 2 kV along with surge current capability. However, the leakage current was as
high as 10-3 A/cm2 at ~ 750 V and reached 10 A/cm2 at 2000 V which needs to be further improved.
Ion implantation methods have also been tried to achieve JBS structures in GaN. First reported in
2016 by Koehler et. al [272], the p-doped regions were formed by Mg ion implantation followed
by symmetrical multi-cycle rapid thermal annealing (MRTA) for activation. The JBS action was
confirmed from reverse bias measurements, which presented much improved leakage current and
BV as compared to a normal SBD. Shortly after, Zhang et al. demonstrated vertical GaN JBS
diodes [270] by (a) Mg implantation into n-GaN to form p-wells and (b) Si implantation into p-
GaN to form n-wells (Figure 52). The implantation and the activation scheme was similar to that
used by Koehler et al. Ron,sp values of 1.5-2.5 mΩcm2 and 7-9 mΩcm2 were observed for the Mg
implanted and Si implanted JBS diodes. Both sets of devices provided 100x-reduction in reverse
leakage current at high reverse bias and presented a BV of 500-600 V.
Trench Metal Barrier Schottky (TMBS) diodes first demonstrated in GaN by Zhang et al. in 2016
[273] represents a device topology which can provide a better control of the reverse leakage current
in an SBD. A TMBS diode consists of a trench metal insulator semiconductor (MIS) structure as
shown in Figure 53(a). The MIS structure does not contribute to the forward conduction phase but
in the reverse bias condition, the two adjacent MIS structures deplete the semiconductor region
between them, thus reducing the leakage current and improving the BV. However, since the
Schottky contact is formed only in a portion of the anode, the Ron,sp is higher than in a conventional
SBD with Schottky contact in the entire anode area. The leakage current can be controlled by
optimizing the trench depth and the TMBS pillar width (Figure 53(b)). But too deep a trench will
also result in premature breakdown of the SiNx dielectric layer due to electric field peaking at the
trench bottom corner. To address this issue, argon-implanted field rings were employed to protect
the base of the TMBS anode. The TMBS diode improved the leakage current by 104-fold and
improved the BV from 400 to 700 V.
(a) (b)
Thus, high performance vertical SBDs with low Ron,sp and high BV have been demonstrated in
bulk GaN and sapphire substrates by improving the material quality of the GaN epitaxial layers as
well as by employing various leakage current mitigation and edge termination schemes. Current
collapse free operation with zero reverse recovery characteristics could be achieved, thus making
GaN SBDs an ideal candidate for low loss rectification purposes.
The first GaN-based CAVET structure for high voltage applications were proposed by Yaacov et.
al in 2004 [276]. The CAVET structure is similar to double-diffused MOS [277] (DDMOS)
structure and comprises of an AlGaN/GaN heterostructure at the top, current blocking layers
(CBLs) and a n-type doped GaN layer at the bottom as shown in Figure 54(a). The source terminals
form ohmic contact to the 2DEG and the gate forms Schottky contact to AlGaN. The CBL
implemented using Mg-ion implantation restricts the flow of current to a small aperture region
which sits just below the gate. By applying a gate bias, the 2DEG below the gate can be switched
on/off, thus resulting in a transistor behavior. The main foreseen advantage is that under voltage
blocking condition, the high field region would sit under the gate in the bulk of the device, unlike
a lateral HEMT, and thus may support large BV as surface related breakdown was eliminated.
However, CAVETs are generally normally-ON devices, as they rely on an AlGaN/GaN channel
which is modulated by the gate. Chowdhury et al., in 2008 [274] demonstrated the first E-mode
CAVET with a threshold voltage of 0.6 V achieved by CF4 treatment in the gate region prior to
gate metallization. Since the AlGaN/GaN layers are regrown by MOCVD after the formation of
CBL, the device threshold voltage varied considerably due to Mg diffusion to the regrown layers.
Chowdhury et.al, in 2012 [278] demonstrated a GaN CAVET with MBE regrown AlGaN/GaN
layers which provided a low Ron,sp of 2.2 mΩcm2 and a BV of 200-260 V at a VGS of -15 V. The
Ron,sp was reduced to 0.4 mΩcm2 [245] by using a buried conductive p-GaN layer as the CBL.
Avogy developed on this idea and in 2014 demonstrated a modified version [279] of these CAVET
structures with a p-GaN gate layer between the gate electrode and the AlGaN barrier for normally-
off operation. Their device provided a forward current as high as 2.3 A and a threshold voltage of
0.5 V. A high BV of 1.5 kV was observed at a VGS of -5 V aided by implanted edge termination
structures. Panasonic corporation [275] demonstrated a similar device by placing a portion of the
channel on the sidewall of an etched trench (forming the gate) and a p-GaN gate (Figure 54 (b)).
They could achieve a low Ron,sp of 1 mΩcm2 and an excellent BV of 1.7 kV (Figure 55). Ji et al.
[280] demonstrated a GaN CAVET with a similar gate trench structure but without the p-GaN gate
and achieved 20 V threshold voltage, and a BV of 225 V. A higher BV of 880 V was achieved by
Ji et al. [281] in a subsequent report by improving the gate trench etching quality and by using a
gate dielectric, which reduced the gate to drain leakage. However, as a result of using a MIS
structure for the gate, the VGS shifted to -21 V from 20 V for the previous report.
Some aspects associated with the GaN CAVET need to be addressed to make this transistor viable
for commercial use. The main issue stems from the regrowth process, which introduces high source
to drain and gate to drain leakage currents in the majority of the reported devices. The on-state
performance is also very sensitive to the doping and dimension of the aperture region formed
between the adjacent CBLs, which is difficult to control due to the relatively complicated
fabrication process.
GaN trench MOSFETs were first reported by ROHM Co. Ltd. in 2007 [284]. These MOSFETs
were fabricated on GaN layers grown by MOCVD on sapphire substrates. Two different gate
dielectrics were investigated: electron-cyclotron-resonance (ECR) deposited SiO2/SixNy and
PECVD SiO2. A reduction in threshold voltage from 25.5 V to 5.1 V was observed by using the
ECR deposited dielectric pair. This work also reported on a high channel mobility of 133 cm2/Vs.
In 2008, the same group demonstrated the first fully-vertical MOSFET [285] on bulk GaN
substrates with similar performance figures as the previous report. Kodama et al. from Toyota
Central R&D Labs [286] devised a method of achieving a smooth m-plane trench sidewall by dry
etching followed by tetra methyl ammonium hydroxide (TMAH) wet etching. TMAH etches m-
plane and c-plane of GaN at a much slower rate than the other facets. Hence sufficient treatment
of the dry etched trench sidewall in heated TMAH results in a smooth surface, which is
predominantly m-plane. This is very beneficial for achieving smooth vertical sidewalls, which
could improve the channel mobility of the MOSFETs. Improvements in BV and Ron,sp of GaN
trench MOSFET were reported by Toyoda Gosei corporation in 2014 [287] (Figure 56 (a)). They
demonstrated a GaN trench MOSFET with a field plate termination achieving a BV of over 1.6
kV and Ron,sp of 12.1 mΩcm2. In 2015 [282], they reported a similar device with Ron,sp improved
to 1.8 mΩcm2, along with a high BV of 1.2 kV. This was achieved by tuning the doping and
thickness of the p-GaN channel region and the i-GaN layer. The Mg doping of the p-GaN channel
layer was reduced, which resulted in lower scattering of inversion channel electrons by the dopant
atoms, and they also slightly increased the doping of the i-GaN layer. Large area trench MOSFETs
[283] with on-state current over 10 A while still maintaining a high BV of over 1.2 kV was also
reported by the same group in 2016, which indicates that dislocation densities from bulk GaN
substrates do not necessarily become a bottleneck for obtaining both high BV and high on-state
current with large area devices (Figure 56(b)). A hexagonal cell array was employed for achieving
a high gate width per unit cell area, leading to an increase in a current density and, thus, a reduction
in the Ron,sp. Very recently [288], their group also demonstrated a vertical trench MOSFET with a
current distribution layer (CDL) below the p-GaN as shown in Figure 57 (a). The CDL consists of
a thin slightly n-type doped (2 × 1016/cm3) layer which can better distribute the current from the
base of the gate trench. This resulted in 1.17x-higher forward current density and an absolute value
of current of 100 A was achieved for large area MOSFETs using the CDL (Figure 57 (b)). HRL
Laboratories developed a method to avoid plasma etch damage to the p-GaN body contact region
by selective area regrowth of n-GaN on top of the p-GaN. They demonstrated a 0.5 mm2 large area
trench MOSFET with an Ron,sp of 8.5 mΩcm2, threshold voltage of 4.8 V and a BV of 600 V. A
detailed analysis of the dependence of main device parameters on gate dielectric thickness, body
layer doping level and cleaning process was presented recently in [289].
(a) (b)
Figure 57 (a) Schematic cross section of a GaN trench MOSFET with CDL
(https://ieeexplore.ieee.org/abstract/document/8757621) [288] (b)Transfer characteristics
measured at VDS of 0.5 V for the MOSFET with and without the CDL
(https://ieeexplore.ieee.org/abstract/document/8757621).
(https://ieeexplore.ieee.org/abstract/document/7520877) [283], [288].
In-situ oxide GaN interlayer-based vertical trench MOSFET (OGFET):
(a) (b)
In a traditional vertical trench gate MOSFET with a n-p-i-n heterostructure, the channel region is
formed by the n-p-i sidewall of the trench gate structure; on the application of a positive gate bias
above the threshold voltage, an inversion sheet charge of electrons is formed in the p-GaN layer
adjacent to the gate dielectric. Since the channel is formed by dry etching, the field effect mobility
of the inversion channel electrons is degraded by the defects in the sidewall formed during the
etching process. In order to alleviate this issue, Prof. Mishra’s group at UCSB devised a technique
[292] whereby a thin undoped GaN layer is regrown in the gate trench region by MOCVD,
followed by in-situ Al2O3 dielectric deposition as shown in Figure 58(a). The initial reports on
these devices presented 60% reduction in Ron,sp as compared to traditional trench gate MOSFETs.
A normally-off operation with a threshold voltage of 2 V was also achieved along with a BV of
195 V. Subsequent devices [290] on bulk GaN substrates provided a higher BV of 990 V aided by
a low damage gate trench etching process. Ji et al. reported large-area (0.2 mm2) OGFETs with an
output current close to 0.5 A and a BV of 320 V. Further optimizations on the growth, design and
fabrication of these devices resulted in OGFETs presenting a record channel mobility of 185
cm2/Vs and a low Ron,sp of 2.2 mΩcm2 [291]. The adoption of a novel double field plate design
helped achieving a high BV of 1.4 kV for a single device and 0.9 kV for a large-area (0.2 mm2)
device (Figure 58(b)). Variations of the OG-FET have been reported by other groups as well. Li
et. al. reported [293], [294] on an trench MOSFET similar to the OGFET but with an MBE regrown
channel rather than by MOCVD. The main aim was to improve on the issue of re-passivation of
p-GaN during regrowth of the interlayer by MOCVD. They could achieve a record high BV of
600 V among GaN transistors with a MBE regrown channel along with reduced thermal budget.
(a) (b)
In recent years, vertical fin power FETs [296] have been demonstrated, with sub-micron fins on
bulk GaN substrates. The advantage over classical trench MOSFETs is that they do not require a
p-GaN layer to provide normally-off operation and blocking under off-state. The gate region of
these devices consists of dielectric/gate metal on the fin sidewalls, which deplete the charge
carriers in the fin due to the work function difference between the gate metal and the GaN,
providing a normally-off operation (Figure 61). [297]
Figure 60 (a) Measured device junction capacitances Cds, Cgs and Cgd .(b) Schematic of the
various Cgs and Cgd components of the FinFET (c) Capacitance component break-out of the
measured Cgs and Cgd. https://ieeexplore.ieee.org/document/8528328 [298]
Figure 61 (a) Structure of the GaN VFET. (b) Electron density simulation in the n-GaN channel
(along the A-A′cut) under different gate bias: for VGS=0V the device is in OFF condition, hence
the electron density in the channel is low and the maximum is located in the center of the n-GaN
region far from the interfaces. At low gate voltage (VGS=1V), the e-density is still relatively low
(≤1015cm4). At high gate voltages (VGS>2 V), the electron density peaks at the Al2O3/GaN interface
[https://ieeexplore.ieee.org/stamp/stamp.jsp?tp=&arnumber=7968303 ] [297]
However, the fin needs to be sufficiently narrow (< 500 nm) to be completely depleted, which
degrades the current capability of the device. The initial devices presented by Sun et al. [296] from
MIT revealed a threshold of 1 V, Ron,sp of 0.36 mΩcm2, and a BV of 800 V with a fin width of 450
nm. The device works by accumulation of electrons rather than by inversion as in the case of trench
gate MOSFETs and high electron mobility of 150 cm2/Vs was obtained in the accumulation layer.
Zhang et al. [295] further optimized the fin width to obtain a higher BV of 1200 V and a lower
Ron,sp of 0.2 mΩcm2 , normalized to the total device area (Figure 59(b)). Large area devices with a
current capability of 10 A and a BV of 800 V were also demonstrated simultaneously. Switching
characteristics [298] of these fin power FETs with a BV of 1200 V and output current capability
of 5 A were compared against commercial 0.9-1.2 kV class Si and SiC power transistors, revealing
the lowest input capacitance (CISS), output capacitance (COSS), gate charge (QG), gate to drain
charge (QGD), and reverse recovery charge (Qrr). These devices exhibited high-frequency (∼MHz)
switching capabilities and superior switching figure of merits (FOMs) as compared to Si and SiC
devices used for comparison. However, these devices break catastrophically, which is possibly due
to the absence of p-GaN layers to modulate the electric field peaks. Also, the fabrication and
control of the fin width could increase the fabrication complexity compared to other vertical
devices. Since the threshold voltage is relatively low ~ 1V, a well-designed gate driver is required
to ensure fail-safe operation. Recently, fin-FET and nanowire-based structures based on a npn
vertical stack have been proposed to obtain robust normally-off operation [299], [300]: the
integration of a p-type layer in a 3D stack (either nanowire- or fin-based) was demonstrated to be
a good strategy for achieving robust normally-off operation, also under gate-stress experiments
[301] (Figure 62).
Figure 62: Cross section of a nanowire GaN device processed on a sapphire substrate. (a) The n-channel
device (Gen1) consists of a 2.5 µm-GaN buffer layer, a 2 µm-GaN channel layer, a 0.5 µm-GaN top layer,
and a 20 nm-SiO2 gate dielectric. (b) The p-channel device (Gen2) comprises a 2.5 µm-GaN buffer layer,
a 0.5 µm p-GaN channel layer, 0.73 µm -GaN and 0.5 µm-GaN as the top layer, and 25 nm-Al2O3 as the
gate dielectric. (c) SEM image of a nanowire of the n-channel device(Gen1); (d) SEM images of a nanowire
of the p-channel device (Gen2) and bird’s-eye view of vertically aligned n-p-nGaN nanowire (NW) arrays
with top contacts [301].
(a) (b)
As presented in the previous sections, the ideal solution for obtaining high quality GaN layers with
defect density less than 106/cm2 would be homo-epitaxy i.e., GaN grown on bulk GaN, as there
would not be any lattice mismatch between substrate and epitaxial layer. However, even after the
demonstration of high-performance diodes and transistors with excellent ON- as well as OFF-state
characteristics, the commercialization of vertical GaN power devices have been hindered by the
high cost and the small diameter of these bulk GaN substrates. Currently, these expensive
substrates are being used only for specific applications in lasers and LED [221]. Hence, in order
to take advantage of the material benefits that GaN offers for power device applications, further
improvements in wafer size and reduction in cost are highly desirable. A strategy to tackle this
issue is by adopting GaN grown on cheaper foreign substrates like Si and sapphire. GaN-on-Si
growth has been widely researched and commercialized for the lateral GaN HEMT technology. A
similar approach could be embraced for vertical power devices as well. Furthermore, GaN-on-Si
growth could give 10-100 times lower wafer + epitaxy cost [219], [303] as compared to bulk GaN.
Silicon substrates also provide better thermal and electrical conductivity as compared to sapphire,
in addition to more mature fabrication processes for the back-end processing. But the main
advantage of GaN on Si is that Si substrates are commercially available up to 12-inch diameters
which could drastically reduce the overall cost per unit of the device. The adoption of Si substrates
could also allow current CMOS compatible fabs to mass produce GaN-on-Si device thus saving
the high cost normally required for setting up new technology fabs. Recently, a new class of
engineered substrates with poly-AlN has been introduced [304]–[307]. The main advantage of
these substrates is their coefficient of thermal expansion (CTE) matched to GaN and thus enabling
the growth of thick, high-quality stress-free GaN with lower defect density as compared to GaN-
on-Si substrates [30, 102]. Thus, the future for GaN-on-Si vertical devices seems very promising.
(a) (b)
Figure 65 Forward and (b) reverse I-V characteristics of vertical PiN diodes. Inset shows the anode
region after destructive breakdown. https://ieeexplore.ieee.org/document/7444154 [308]
                         (a)                         (b)
Figure 66 (a) Schematic cross section of the a fully vertical GaN p-i-n diode on Si substrate grown
by MOCVD (https://iopscience.iop.org/article/10.7567/APEX.9.111005/meta) [310]. (b) Fully
vertical     GaN-on-Si         p-in     diode       by      selective      substrate      removal.
(https://ieeexplore.ieee.org/abstract/document/8325453) [311].
GaN-on-Si p-i-n diodes have been demonstrated since 2014. Zhang et al. demonstrated the first
GaN-on-Si p-i-n and Schottky diodes [302] with a BV of 300 V and 205 V respectively (Figure
63). Even with a high defect density of 109/cm3, a peak electric field of 2.9 MV/cm could be
achieved. The current leakage paths, leakage mechanisms and methods to improve the leakage
current were thoroughly researched and identified in a later work [303]. A device termination
scheme based on ion implantation and anode field plate was found to reduce the leakage current
by almost two orders of magnitude. For GaN layers grown on Si substrates, the traditional device
structure is quasi-vertical, in which both top and bottom layers are accessed through the device top
surface. This results in a non-uniform distribution of current from the anode and current crowding
at the bottom GaN layer near the cathode terminal. In order to alleviate this issue, Zou. et al.,
demonstrated a method of making a fully-vertical p-i-n diode [308] by removing the Si substrate
below the GaN layers, followed by transfer of GaN epilayers to a carrier wafer as shown in Figure
64(a). The finished p-i-n diode presented an Ron,sp of 3.3 mΩcm2 and a BV of 350 V (Figure 65).
Excellent temperature stability up to 175 °C was also observed during ON- as well as OFF-state
measurements. Zhang et al. demonstrated a similar method of achieving fully-vertical p-i-n diodes
[309] but with better Ron,sp of 1 mΩcm2 and BV of 500 V (Figure 64(b)). A low reverse recovery
time of 50 ns comparable to bulk GaN diodes were extracted along with excellent thermal stability
of the devices up till 300 °C. Mase et al. [310], demonstrated a novel method of achieving a fully
vertical GaN-on-Si p-i-n diode by using a conductive n-type Si substrate along with n-type buffer
layers (Figure 66(a)). An interesting feature of their p-i-n heterostructure was the use of a 3.2 µm
thick strained super lattice (SLS) of n-GaN/n-AlN. Sufficiently thick SLS layer (~ 3 µm) was
necessary to control the edge dislocations to an appreciable value ~ 2×109 /cm2 which also dictated
the lowest doping level possible for the GaN layers grown over the SLS layer. The devices
presented an Ron,sp of 7.4 mΩcm2 and a BV of 288 V. A detailed analysis of current crowding
effect in quasi-vertical structures were provided by Zhang et al. in 2017 [205]. According to this
research, the thickness and the doping of the bottom n-GaN current collecting layer determines the
Ron,sp and a thick highly doped bottom n-GaN is required to ensure lower levels of current
crowding. High voltage GaN-on-Si p-i-n diodes were demonstrated by Khadar et al. in 2018 [312].
These diodes had a BV of 820 V with just a 4 µm thick drift layer and an ultra-low Ron,sp of 0.33
mΩcm2 resulting in a record value of 2 GW/cm2 for the BFOM. The growth of these layers was
optimized to obtain i-GaN layers with an excellent electron mobility of 720 cm2/Vs and a low
defect density of 2 × 108/cm3 for GaN grown on Si. Shortly after, Zhang et al. [311] demonstrated
a new method of achieving fully-vertical operation for GaN-on-Si p-i-n diodes by selective Si
removal underneath the active area of the device followed by metallization (Figure 66(b)). An
Ron,sp of 0.35 mΩcm2 and a BV of 720 V was achieved.
(a) (b)
Figure 67 (a) Schematic cross section of the a quasi-vertical GaN power MOSFET on Si substrate
(https://ieeexplore.ieee.org/abstract/document/8125763) [313] (b) Fully-vertical GaN-on-Si
power            MOSFET               by           selective           substrate          removal.
(https://ieeexplore.ieee.org/abstract/document/8621009) [314].
Figure 68 (a) Comparison of the IDS -VDS of the fabricated vertical MOSFETs with gate trench
aligned along m- and a-plane, using metal and oxide hard masks. SEM images of the trench
sidewall aligned along the (b) a-plane and (c) m-plane, after TMAH wet treatment. Notice the
much       smoother        m-plane         sidewalls     compared         to       the     a-plane.
https://ieeexplore.ieee.org/document/8621009 [314]
GaN-on-Si vertical power MOSFETs have been also demonstrated, with the first demonstration
being from Liu et al. in 2017 [313] (Figure 67(a)). Normally-off operation with a threshold voltage
of 6.3 V, which is ideal for power converter applications, along with an ON/OFF ratio of over 108
were achieved. The device presented a low Ron,sp of 6.8 mΩcm2 and a BV of 645 V which is
comparable to bulk GaN power MOSFETs. The first demonstration of fully-vertical MOSFETs
was subsequently devised by Khadar et al. [314] (Figure 67(b)). The method involved a robust
fabrication process including selective substrate removal underneath the MOSFET followed by
ohmic contact deposition and copper electroplating to provide strength to the thin free standing
epitaxial GaN layers. An Ron,sp of 5 mΩcm2 and a BV of 520 V was obtained. These devices
exhibited 2.8x-higher current density and 3x-lower Ron,sp as compared to quasi-vertical control
power MOSFETs, due to absence of current crowding. Also, insights into the impact of hard mask
selection for gate trench etching and gate trench alignment to either the m- or a-plane sidewall, on
the output current density was analyzed (Figure 68). Devices having gate trench aligned along the
m-plane provided 3x-higher output current as opposed to those with gate trench aligned along the
a-plane. The use of a metal mask for gate trench etching provided a high inversion channel field-
effect mobility of 41 cm2/Vs for electrons.
GaN-on-Si technology offers a unique advantage for the possible integration of several different
devices on the same chip to realize integrated circuits (IC) which have evident benefits like smaller
IC foot print, greatly reduced parasitic capacitance and resistance arising from wire bonding of
discrete devices leading to higher efficiency, lower cost, etc. To date several different integration
schemes on GaN-on-Si lateral technology has been demonstrated with HEMTs [315]–[318]. Liu
et al. demonstrated the first vertical monolithically integrated device [319] in 2018, where a
freewheeling Schottky barrier diode was integrated with the power MOSFET (Figure 69), to
overcome the lossy body diode by using a fast low turn-on voltage SBD. In several topologies of
power converters, such as buck/boost converters, voltage-source inverters, and resonant
converters, where an inductive load is controlled by switches, a freewheeling diode parallel to the
power MOSFET is required to allow a reverse flow of current when the supply current to the load
is suddenly interrupted. The integrated SBD was created by dry etching the top n- and p-GaN
layers followed by TMAH treatment to smoothen the surface and then Schottky metallization. The
integrated MOSFET/diode provided excellent forward and reverse characteristics. In particular,
the freewheeling diode presented a low turn-on voltage of 0.76 V, low Ron,sp of 1.6 mΩcm2, ideality
factor of 1.5 and an excellent BV of 254 V achieved without any additional termination
mechanisms.
Figure 69 (a) Equivalent circuit, (b) Schematic of integrated vertical MOSFET-Schottky barrier
diode (SBD). (c) SEM image of integrated vertical MOSFET-SBD, (d) Cross-sectional SEM
image of the integrated vertical MOSFET, and (e) of the integrated vertical SBD.
(https://ieeexplore.ieee.org/abstract/document/8370664) [319]
Recently, the optimization of GaN-on-Si semivertical devices has been the subject of pioneering
projects in the field, working with silicon substrates up to 200 mm in diameter. As an example, we
mention recent papers aimed at optimizing the full GaN-on-Si structure, with focus on the Mg-
doping in the p-body layer [289], the thickness of the gate insulator [289], and the dielectric
material used at the gate [320], with the aim of minimizing on-state resistance and charge trapping,
and of obtaining high breakdown voltage [321].
GaN based vertical power devices could be used for realizing the next generation power efficient
converters provided that pertinent issues related to material quality, device fabrication and
performance can be solved. Compared to Si and SiC, bulk GaN crystal growth process is highly
challenging requiring high temperatures > 2200 °C and nitrogen pressure > 6 GPa making it
impossible to crystallize GaN from the melted compound [211]. Hence, lower pressure and
temperature methods like hydride vapor phase epitaxy (HVPE), sodium flux and acidic/basic
ammonothermal growth method has been utilized to realize bulk GaN substrates. Even so, the
defect density in these substrates are around 104-106 /cm2, the effect of which has to be closely
scrutinized for use in commercial applications. In general, Si and SiC devices are qualified by the
fully Joint Electron Device Engineering Council (JEDEC), which will also have to be done for
GaN-based vertical power devices. For GaN HEMTs it was realized that conventional JEDEC
guidelines for Si based devices are not sufficient and led to the creation of new JEDEC standards
JC-70, JEP-180 and JEP-173 for GaN based devices [322]–[324]. A similar situation may arise for
vertical GaN, and guidelines used for Si, SiC and lateral GaN may be adopted or refined, following
a in depth discussion by the involved community.
Conventional Si and SiC power devices rely on the selective doping capability by ion implantation
or by diffusion to achieve junction termination structures (JTEs) which have been proven to be
indispensable for achieving reliable devices with avalanche capability [325]. However, for GaN,
doping by ion implantation is still a highly complex process requiring specialized instruments
capable of high-temperature [326] (> 1200 °C) and high-pressure conditions [327] (> 1 GPa).
Selective area growth is another option which has been investigated but with little success due to
the presence of high concentration (> 1018 /cm3) of Si atoms at the regrowth interface [236], [328].
Several schemes like Ar ion implantation, plasma-based treatments described in the previous
sections, have been adopted for GaN based vertical devices, however their suitability for large
scale production and long-term reliability is still unclear. There are also concerns related to the
low activation efficiency (~ 1%) of p-GaN grown by MOCVD due to the high bond energy of the
stable Mg-H complexes formed during the growth process [329]–[331]. Even high temperature
annealing (~ 750 – 850 °C) activates only a small fraction of the dopant atoms (Mg concentration
> 1019 /cm3) thus presenting a low hole concentration typically around 1017 /cm3 [332], [333] with
low hole mobility due to scattering from the dopant atoms present in large concentrations. This is
in particular not desirable for GaN trench power MOSFETs since the high concentration of dopant
atoms in the p-GaN channel region scatters the electrons formed by inversion during normal ON-
state operation, thereby degrading the channel mobility [290], [313], [334] and thus the output
current. The low hole mobility of the p-GaN layers also affects the resistance of p-i-n diodes and
results in ON-state losses.
Adoption of bulk GaN for commercialization is not only hindered by the high cost and the small
size of these substrates but also by the huge initial investment needed for the setting up GaN
specific fabs, which could shoot up the average selling price (ASP) of discrete devices slowing
down further the adoption of these devices. Hence it is vital to make significant strides in the
improvement of substrate size as well as cost of these bulk GaN substrates.
GaN-on-Si could offer an alternative to lower the device costs, but significant progress has to be
made on the GaN quality with low defect density and on the thickness of the GaN layers grown
on Si substrates, in order to enable larger voltage devices. The advent of CTE matched substrates
available in large wafer diameters [335] may further push the development of vertical GaN power
devices.
8 Charge-trapping processes in GaN transistors
The presence of deep-level traps [336]–[338] in GaN and its alloys and the effects of the associated
charge capture/emission processes can rarely be neglected in GaN transistors under their typical
operating conditions in power RF amplifiers and power switching converters. As will be detailed,
figures of merit that are key to these applications and that can detrimentally be affected by trapping
processes include output power (Pout) and power-added efficiency (PAE) in RF amplifiers,
switching and conduction losses in power converters.
Spatial position of traps within the device (surface, barrier, buffer and, if present, gate dielectric),
trap parameters [337], [338] (energy, capture cross sections, concentration, acceptor- or donor-like
behavior), type of involved carriers (electrons or holes) and the associated charging/discharging
path (contact injection/removal, carrier generation, internal redistribution) can all play a role, and
their signatures on transistor operation should be known by the technologist for a proper device
optimization.
In addition, several trap measurement methods have been devised and extensively applied to GaN
structures and transistors, in order to quantitatively characterize traps.
This section is organized as follows. First, the properties of the most common traps in GaN are
reviewed, to help the reader understanding the complexity of the topic. Second, the most important
trapping mechanisms affecting GaN transistors are elucidated by crossing the different trap
locations with the corresponding possible charging/discharging paths. Then, the resulting trap
effects and their relevance for applications are reviewed. Finally, methods that can be adopted for
measuring trap parameters are reviewed.
8.1     Traps and deep levels in GaN
The properties of gallium nitride as a semiconductor material are determined by the specific
periodicity of its crystalline structure. If this periodicity is lost, either due to incorporation of
foreign atoms into the crystal or to a non-ideal arrangement of the host atoms during the growth
of the material, a defect is locally generated. Defects are referred to as “impurity-related” in the
former case, whereas they are defined as “native” in the latter. The presence of these defects locally
perturbs the periodic potential of the lattice, therefore introducing allowed energy states within the
forbidden bandgap of the material. These states, that can either act as carrier traps or recombination
centers, can have a detrimental impact on both the performance and the reliability of GaN-based
devices. It is therefore of primary importance to identify the properties and the physical origin that
such states have in actual GaN devices. To this aim, a database of traps and deep-levels related to
the GaN material system has been created by collecting and comparing the data of almost one
hundred scientific publications (Table 7). The comparison of the 480+ records of traps
experimentally detected in GaN by means of various techniques allowed us to correlate energy
positioning, signature and trapping behavior of the levels accredited to the defects most commonly
found in GaN. In the following paragraphs we report the outcome of the study of such defects
based on energetic-positioning considerations.
   Type of
                                           Reported Ea         Detected
  defect or       Physical origin                                                 References
                                              (eV)                as
  impurity
                                        NATIVE DEFECTS
                                                                             [357]–[361], [362]*,&,
   Nitrogen                                  1 to 0.76            EC -
                     NI-related                                                      [363]
  interstitial
                                               2.42              EV +                [364]
Native defects
The position of the trap levels associated with native defects in GaN are reported in Figure 70.
These levels are mainly related to nitrogen interstitial (NI), nitrogen antisites (NGa), nitrogen
vacancies (VN), gallium interstitials (GaI), gallium vacancies (VGa), to clusters of the former or to
unspecified extended defects (mostly dislocations).
Nitrogen interstitials have been reported to introduce trap states between EC - 1.2 eV and EC –
0.76 eV [357]–[364], with average preferential energy positioning at EC – 1.02 eV, EC – 0.89 eV
and EC - 0.79 eV. The former and the latter levels have also been tentatively associated with an
extended defect in [362]. Among nitrogen-related native defects, nitrogen interstitials exhibit the
highest activation energy, allowing them to act both as traps or recombination centers.
            EC-------
                3.4
                          3.2
            Energy (eV)
3.0
2.8
2.6
2.4
                0.9
                0.6
                0.3
            EV-------
                                NI   NGa      VN       GaI      VGa Extended
 Figure 70 Energy states detected in GaN and associated to native or extended defects. The
 reference level, assumed to have an energy equal to 0, is related to the top of the valence band.
 The scatter-like representation highlights the preferential energy assignment of the detected
 traps with respect to a specific type of defects.
Nitrogen antisite defects are assumed to form a mini-band of levels between EC – 0.66 eV and EC
– 0.5 eV [358], [362], [365]–[375]. None of the aforementioned reports suggests a direct
correlation between the levels associated to nitrogen antisite defects and extended defects.
Nitrogen vacancies represent the most commonly observed type of native defect in GaN, with
more than 40 records in 25 different scientific publications [358], [362], [365]–[375]. The
associated trap states are preferentially positioned in the EC – 0.27 eV to EC – 0.089 eV range, with
most of the occurrences belonging to the mini-band at EC – (0.24 ± 0.003) eV identifying the
typical level associated to nitrogen vacancies in GaN. The outliers of the aforementioned
distribution are represented by the deeper EC – 0.35 eV [359], EC – 0.4 eV [386], EC – 0.53 eV
[387] and EC – 0.613 eV [379] levels, respectively associated with VN complexes, simple nitrogen
vacancies, triply ionized nitrogen vacancies and clusters of vacancies. Other VGa-related levels
whose physical origin was found to be compatible with an extended defect are located at EC – 0.19
eV [368], at EC – 0.23 eV [367], [368] and at EC – 0.25 eV [383]. The fact that trap levels with
similar Arrhenius signatures, and therefore similar activation energy and cross-section, have been
associated both with extended and point defects suggests that even if the capture rate of the defect
is influenced by the electrostatic repulsion due to the close proximity to other ionized traps or to
extended crystal defects, gallium vacancies tend to maintain their characteristics emission
properties.
Gallium interstitials are reported as the physical origin of traps detected in GaN by only a couple
of papers. In [389], an EC – 0.91 eV level was associated with GaI complexes located along
dislocations. In [363], a trap located at EC – 0.8 eV was tentatively ascribed to gallium interstitials
located in an unspecified layer of an AlGaN\GaN HEMT.
Gallium vacancies are often associated with deep levels responding in the EC – 2.42 eV to EC –
2.85 eV range, with most occurrences roughly located at EC – 2.6 eV [360], [363], [364], [366],
[382], [384], [390]–[395]. Some of these traps have been tentatively associated with oxygen-
[360], [363] or hydrogen- [382], [392] related complexes formed with VGa. Also the EC – 1.12 eV,
the EC – 0.64 eV and the EC – 0.6 eV levels found in [371] have been referred to as VGa-O-related
defects, whereas the other deep levels detected outside the band of reference for this particular
type of native defect have been associated to VN-VGa complexes (at EC – 0.24 eV in [365]) or to
simple gallium vacancies (at EC – 0.62 eV in [383], at EC – 1.64 eV in [375] and at EC – 3.19 eV
in [386]).
Trap states related to extended defects have been found to cover a wide and non-continuous
portion of the upper part of the GaN bandgap, ranging from EC – 1.118 eV to EC – 0.24 eV [359],
[362], [366], [369], [371], [377], [384], [388], [396], [397]. Interestingly, the upper group of levels,
scattered from EC – 0.27 eV to EC – 0.17 eV, covers the same range of activation energies exhibited
by VN-related defects. Similarly, also a second mini-band of deep levels associated to extended
defects, and scattered from EC – 0.641 eV to EC – 0.41 eV, finds a correspondence with the typical
range of response of NGa-related defects. These considerations suggest that both nitrogen vacancies
and nitrogen antisite defects can be found in proximity of other defects, and therefore behave as
interacting defects. Interestingly, no defects with activation energies corresponding to the VGa band
were ascribed to extended defects, possibly suggesting that this type of native defect tends to
predominantly behave as a non-interacting point defect, or does not form in highly defective device
regions.
Impurity-related defects
Imperfections of the lattice structure may also arise from the inclusion of atomic species that
nominally do not belong to the GaN crystal. Such atoms, or atom complexes, are referred to as
impurities. These impurities can either be intentionally introduced into the material, as in the case
of dopants, be originated as the byproduct of the chemical reactions occurring during crystal
growth, or derive from unwanted contaminations. Si and Mg represent the main doping species for
GaN. Similarly, Fe and C can be intentionally introduced in order to compensate for the intrinsic
n-type conductivity of GaN, Others may act as unwanted contaminants, like residual O and H
atoms. A more detailed description of the deep levels introduced by these impurities is provided
in the following paragraphs.
            EC-------
                3.4
                          3.2
                          3.0
            Energy (eV)
                          2.8
                          2.6
                          2.4
                          0.9
                          0.6
                          0.3
            EV-------
                                C   H        Fe        Mg         O        Si
 Figure 71 Energy states detected in GaN and associated to impurities-related defects. The
 reference level, assumed to have an energy equal to 0, is represented by the top of the valence
 band. The dispersion of the detected levels within the lower portion of the forbidden gap has
 been magnified with respect to Figure 70.
Aside from its shallow donor level, which has a typical activation energy ranging from EC – 0.011
eV to EC – 0.028 eV for moderately doped GaN layers [414]–[416], only few scientific reports
attribute to silicon deeper allowed energy levels, respectively located at EC – 0.37 eV [377], EC –
0.4 eV [377], EC – 0.59 eV [398].
Magnesium is employed in GaN devices in order to achieve p-type conductivity. Trap states
detected near the valence band edge at EC – 3.2 eV [399], EC – 3.22 eV [382], [393], EC – 3.25 eV
[395] and EC – 3.28 eV [375], even if nominally assigned to other physical origins, can reasonably
be ascribed to the shallow level of the dopant, which is known to be typically located from 0.15
eV to 0.2 eV above the valence band edge [417], [418]. A shallower level at EC – 3.36 eV [400]
was tentatively ascribed to Mg-H complexes, and a similar origin was assigned to the EC – 0.62
eV level in [393]. Additional levels at EC – 0.355 eV [376] and at EC – 0.597 eV [401] were ascribed
to generic Mg-related defects, whereas the EC – 0.44 eV level in [385] was associated with Mg-VN
complexes.
Due to its unpaired valence electron, hydrogen is prone to form complexes with other chemical
species in addition to magnesium. Different authors ascribe deep levels located between EC – 2.62
eV and EC – 2.47 eV to complexes formed between hydrogen and gallium vacancies [382], [392],
even if the similarity between this range and the preferential energy positioning of VGa-related
defects indicates that those levels may be ascribed to the latter. Levels at EC – 0.49 eV and EC –
0.578 eV were tentatively assigned to carbon- and/or hydrogen-related defects, respectively in
[403] and [402].
Due to its very-high electronegativity, also oxygen tends to bond with other impurities of native
defects within the GaN crystal. One of the few literature reports investigating the formation of
deep levels related to O in GaN associates the levels at EC – 1.118 eV, EC – 0.642 eV and EC –
0.599 eV either to VGa-O complexes or to dislocations [371]. In [404] the EC – 0.44 eV level was
ascribed to ON. The same interpretation was provided by the authors in [386] to tentatively explain
the shallow EC – 0.04 eV donor level detected by means of PL measurements as part of a donor-
acceptor pair.
Iron is often adopted in GaN-based transistors to intentionally reduce the intrinsic n-type
conductivity of specific u.i.d. GaN layers. Several levels, ranging from EC – 0.94 eV to EC – 0.34
eV, have been detected and associated to Fe [370], [390], [405]–[408]. In particular, the authors
in both [408] and [407] agree when associating shallower deep states, respectively detected at EC
– 0.34 eV and EC – 0.44 eV, to transitions between, or related to, different charge states of the Fe
atoms. Recent reports deeply investigated the signatures of Fe-related traps [151], [419], [420]; it
was recently suggested that FeGa sites are related to a trap located at EC-0.58 eV, =2x10-15 cm2
[419].
Like iron, also carbon doping is used to compensate the intrinsic n-type conductivity of the GaN
crystal. This effect is achieved when C atoms occupy a substitutional position on the nitrogen site
(CN). CN is a deep acceptor, with the (0/-) transition level theoretically located at EV+0.9 eV [421],
and experimentally associated with a band of allowed states between EC – 2.64 eV and EC – 2.49
eV [365], [366], [400], [411]–[413]. Being a deep acceptor, the CN level does not generate large
free hole density and strong p-type conductivity; large concentrations of CN can pin the Fermi level
at EV+0.9 eV, leading to semi-insulating layers. CN may also induce the formation of shallow
acceptor-like levels in the EC – 3.31 eV to EC – 3.15 eV band [366], [382], [384], [388], [389],
[392]–[394], [400], [409]–[412]. On the other hand, if C occupies interstitial positions, the
resulting CI sites are believed to introduce a deep donor level in the upper part of the gap, with
typical energies ranging from EC – 1.35 eV to EC – 1.2 eV [382], [384], [388], [393]. Other
shallower levels at EC – 0.14 eV [403], EC – 0.4 eV [365], EC – 0.49 eV [403] and EC – 0.578 eV
[402] have been tentatively ascribed to carbon- or hydrogen-related defects, with a speculative
assignment of the EC – 0.4 eV to CGa sites.
8.2   Trapping mechanisms
Traps influence the electrical behavior of transistors since they are characterized by relatively long
capture/emission times [337] and thus they charge up or discharge following fast bias changes
more slowly than device capacitances that govern the “prompt” device response.
Regardless of the nature of traps, i.e. whether they are associated to intrinsic crystallographic
defects, unintentional or intentional impurities or defect-impurity complexes, only those traps that
have a chance to change their charge state during the device functioning through capture or
emission of mobile carriers can induce electrical effects of concern for the dynamic device
performance. Based on these considerations, trapping mechanisms can be categorized as the
combination of a trap location plus an associated charging/discharging path and type of involved
mobile carriers.
Traps can virtually be located in any semiconductor or dielectric layer as well as at interfaces, but
locations that have more frequently been associated with harmful effects for GaN transistors
include the device surface and the passivation within the gate-drain access region, the barrier, the
buffer (including the interface with the barrier layer), and, if present, the gate dielectric (including
the interface with the underlying semiconductor). The reported charging/discharging paths include
all leakage currents from the device terminals (gate, source, drain, substrate), the 2DEG at the
barrier/buffer interface in HEMTs and unrecessed or partially-recessed MIS-HEMTs or the
channel at the dielectric/GaN interface in fully-recessed MIS-HEMTs, charge redistribution
processes within floating C-doped buffer, as well as high-field mechanisms like field-enhanced
emission, trap impact ionization, and Zener trapping. One or more of these paths can be activated
depending on the applied device bias.
Two general remarks that will apply to all effects described below, are as follows.
1) Since all GaN transistors relevant for applications are n-channel field-effect transistors, traps
can induce a reduction in the source-to-drain channel conductivity and, through this, in the drain
current, if the negative trapped charge increases or the positive one decreases. These changes can
either be the result of electron capture or hole emission. On the other hand, an increase in the
channel conductivity and drain current can be promoted by electron emission or hole capture.
2) Trapping effects that take place in the device portion under the gate directly influence V T,
whereas those occurring in the gate-drain access region impact the drain access resistance and the
transconductance peak.
Table 8 lists the major trapping mechanisms that have been reported in the literature classified in
terms of trap location and corresponding charging/discharging path and type of involved mobile
carriers. The different trapping mechanisms resulting from this classification will be described in
more detail in the following sections.
Nucleation/transition Nucleation/transition
      Substrate                                                             Substrate
                                    Sub                                                                   Sub
                                    0V                                                                    0V
                                        (a)                                                                   (b)
      0V                  <<0V                             0V                0V                >>0V                             0V
                      G                                                                         G
       S                          Ox1                       D                S                                                  D
                                                                                                        Ox2
      Barrier                                                               Barrier
      Buffer                                                                Buffer
      Nucleation/transition                                                 Nucleation/transition
      Substrate                                                             Substrate
                                    Sub                                                                   Sub
                                    0V                                                                    0V
Surface traps in the gate-drain (G-D) access region (in following, simply, surface traps) can capture
electrons injected by the gate contact in GaN HEMTs [mechanism Su1 in Table 8]. This occurs as
the device is biased under off-state conditions, i.e. when the gate is biased below threshold, and a
large, positive VDS is applied. Gate-injected electrons can leak across the passivation [339], [422]–
[424] or propagate through surface traps by a hopping mechanism [425], [426]. Trapped electrons
are then slowly emitted by traps following the device switching to on-state conditions, i.e. zero or
positive VGS with small, positive VDS. To describe this mechanism, the “virtual gate” [339]
concept has been used. In the early stage of the GaN HEMT development, this was the trapping
process of major concern owing to the heavy limitation it could induce on the achievable RF output
power and power-added efficiency. Passivation optimization [427], [428] and field-plate
introduction [429], [430] were key to minimize it, thanks to a combination of surface defect density
reduction (passivation optimization) and gate-drain electric field mitigation (field plates).
Another possible charging path for surface traps is through trapping of 2DEG electrons gaining
sufficient energy to overcome the barrier due to the AlGaN/GaN conduction-band offset [199],
[431]–[434] [mechanism Su2 in Table 8]. This process requires the presence of channel hot
electrons (CHE) and therefore an accumulated device operation under semi-on or power-state
conditions (i.e. simultaneously high VGS or ID and VDS), like in RF power or hard-switching
conditions.
As far as the physical nature of surface traps is concerned, dominant discrete levels [55] as well as
distributed interface states [56], [435]–[437] or border traps [438] have been reported.
Finally, since these traps are located within the drain access region, their primary, detrimental
effect is an increase in the drain access resistance and therefore a reduction in the transconductance
peak.
In GaN HEMTs barrier traps located under the gate can capture electrons injected by the gate
contact following off-state biasing and slowly emit them as the device is turned to on-state
conditions [mechanism Ba1 in Table 8]. Under power-state conditions, hot electrons from the
2DEG can instead be trapped by barrier traps located in the G-D access region [432] [mechanism
Ba2 in Table 8]. The consequence is either a positive shift in VT or an increase in the drain access
resistance, depending on whether traps are located under the gate or in the G-D access region.
During drain voltage sweeps, donor-like barrier traps can be ionized by field-enhanced electron
emission [439], [440] [mechanism Ba3 in Table 8], in this case inducing an increase in the output
conductance sometimes termed “kink effect” (see Section 8.7.4).
In normally-off p-GaN HEMTs, hole trapping can take place at the p-GaN/AlGaN interface under
positive gate bias [87] [mechanism Ba4 in Table 8].
Buffer traps are probably the major source of trapping effects in present-days GaN transistors,
following the already mentioned advancements in surface passivation and the introduction of field
plates, which resulted in the effective minimization of surface trapping effects.
Buffer-trap effects have been associated with several different trapping mechanisms. The most
straightforward charging path is through electrons originating, under off-state bias, from the source
and/or gate contacts and forming the source-drain [441]–[443] [mechanism Bu1 in Table 8] and
the gate-drain [343] [mechanism Bu2 in Table 8] leakage currents, respectively. At very large
drain voltages (always under off-state conditions), the substrate leakage current come also into
play, contributing to electron trapping into buffer traps mainly within the G-D access region under
the drain contact [444] [mechanism Bu3 in Table 8]. Under semi-on and power-state bias
conditions, on the other hand, energetic 2DEG electrons can be injected deeply into the buffer and
get thereby trapped, mainly in the G-D access region [434], [445] [mechanism Bu4 in Table 8].
Other buffer trapping mechanisms that have been invoked include the capture of valence-band
electrons through Zener processes involving traps in the buffer region under the gate edges [446]
[mechanism Bu5 in Table 1], and at the barrier/buffer interface [447] [mechanism Bu6 in Table
8]. Moreover, trap impact ionization by channel electrons has been proposed as a mechanism that
can discharge traps at the barrier/buffer interface as the drain bias is increased above some critical
voltage and resulting in the so-called “kink effect” in the device output characteristics [355]
[mechanism Bu7 in Table 8].
With relevance to power transistors with C-doped buffer, peculiar trapping effects taking place in
the weakly p-type buffer characterizing these devices have been explained by the so-called “leaky-
dielectric” model [448] to describe charge injection into the buffer region [mechanism Bu8 in
Table 8] or by hole redistribution within the buffer itself [449] [mechanism Bu9 in Table 8].
As far as the physical origin of buffer traps is concerned, either GaN intrinsic defects [443], [450],
[451] or intentional impurities have been correlated with the observed trapping effects. The latter
are acceptor-like traps purposely introduced to suppress the n-type conductivity of GaN. Fe and C
are the two species that are more commonly adopted.
Fe is generally adopted for RF transistors [443], [452]. The dominant deep level introduced in the
GaN bandgap by Fe doping is generally assumed to be energetically located at 0.5-0.6 eV from EC
[370], [405]. As a result, the Fe-related level behaves as an acceptor-like electron trap.
C is typically used for power devices [67] and can be incorporated (i) by tuning growth parameters
to control the decomposition rate of TMGa and the incorporation of C from methyl group
(autodoping [124], [453]), or (ii) by adding a C precursor (propane, methane, ethylene) [454] to
CVD gas mixture (extrinsic doping). Early DLTS/DLOS measurements [455] and correlation with
DFT calculations based on LDA approximation [456] suggested two deep levels related to C
doping being the CN (acceptor) and CGa (donor) centers, respectively located at EV+0.14 eV (or
equivalently EC-3.28 eV) and EC-0.11 eV. More accurate and recent hybrid-functional DFT
calculations yielded CN at EV+0.9 eV and CGa above EC, along with other C-related centers inside
the bandgap like interstitial CI center at EC-0.4 eV (acceptor) and the CGa-VN complex at EC-0.1
eV (donor [421], [457], [458]). While there is nowadays a rather general consensus that trapping
effects in C-doped GaN transistors are mainly related to the dominant CN traps, many simulation-
based works [152], [449], [459]–[464] suggest that some degree of auto-compensation between
these acceptors and C-related donor traps must take place, reducing the effective acceptor density
below the level of the introduced C concentration (especially in case of extrinsic C doping). On
the other hand, higher donor concentration in GaN:C compared with donor density in undoped
samples has been confirmed also experimentally [465]. The (EV+0.9 eV) level is anyway able to
induce the compensation of the buffer region, for suitable concentrations. As mentioned above,
this may lead to peculiar trapping effects, which are still subject of investigation in the literature
[460], [466]–[469].
Regardless of whether they are related to intrinsic defects or intentional impurities, buffer traps are
essential to compensate the unintentional n-type conductivity (related to VN, Si, O) of nitrides,
therefore allowing for an abrupt channel pinch-off, reducing the source-drain leakage current and
increasing the off-state breakdown voltage. Owing to this, buffer optimization requires a trade-off
between trapping effects and breakdown voltage. In the case of doped buffers, this is typically
achieved by switching off the impurity flow during growth at a designed distance from the channel
at the barrier/buffer interface [442], [470].
Similarly to barrier traps, buffer electron traps can lead to a positive shift in VT or an increase in
the drain access resistance and RON depending on whether involved traps are located under the
gate or in the G-D access region. Positive charge can also be accumulated in the buffer, leading to
a decrease in RON, under specific conditions [448], [471].
In transistors with an insulated-gate structure, like AlGaN/GaN MIS-HEMTs, also the gate oxide
bulk and its interface with the underlying semiconducting region can act as trapping sites.
As far as bulk oxide traps are concerned, the main charging/discharging path is through the gate
electrode [mechanism Ox1 in Table 8], which can inject electrons into the gate dielectric at large
and negative VGS [472], [473] and remove trapped electrons out of the device under positive VGS
[474].
For interface and border traps, the main charging path is through electron injection from the device
channel at the AlGaN/GaN interface (in MIS-HEMTs) or the GaN surface (in MOSFETs)
[mechanism Ox2 in Table 8. Classification of major trapping mechanisms in terms of trap location,
charging/discharging path and type of involved mobile carriers. Table 8], that can be triggered at
positive VGS (forward-gate bias [349]). These traps can instead emit electrons at negative VGS
[353].
Another reported mechanism, specific to devices with a C-doped buffer, consists of hole capture
into interface traps, where holes are provided by C-related acceptors in the buffer and not
necessarily by a high-field electron-hole generation mechanism [mechanism Ox3 in Table 1
[464]].
Both bulk and interface dielectric traps affect the device performance through VT, by inducing
instabilities on this crucial parameter [475]. These effects will be described in more detail in
Section 8.7.3.
In this section the major charge trapping effects observed in GaN transistors are reviewed and
associated to the mechanism(s) put into evidence in the previous section. The correlation between
trapping effects and related mechanisms is summarized in Table 9.
Table 9. Major trapping effects in GaN transistors with associated, responsible mechanism(s)
(labelled according to Table 1).
        Trapping effect                       Trapping mechanism(s)
Dynamic RON Su1, Su2, Ba1, Ba2, Bu1, Bu2, Bu3, Bu4, Bu5, Bu6, Bu8, Bu9
Responsible mechanisms are related to electron trapping taking place under off-state or semi-on
bias. Channel conductivity is weakened by this and can not promptly be restored as the dynamic
operating point moves to ON-state. Electrons can be provided by the gate through tunneling
injection at large drain-gate voltage, by the source via source-drain leakage current due to the large
drain-source voltage, or by the 2DEG channel under semi-on conditions. Gate electrons can reach
surface, barrier, as well as buffer traps, whereas source electrons can only get trapped into buffer
traps. Channel hot electrons can be trapped by buffer, barrier or surface traps. Either V T or the
drain access resistance or both parameters can in principle be increased, depending on whether
electron trapping takes place under the gate, in the gate-drain access region or both. A major impact
is typically associated to electron trapping within the gate-drain access region. Optimization of
surface passivation [427], [428] and introduction of field plates [429], [430] have comparatively
decreased the role of surface and barrier traps as promoters of the RF current collapse, thanks to
the mitigation of the electric field at the drain-end edge of the gate, leaving buffer electron trapping
as the major responsible mechanisms, especially in technologies where compensating impurities
(like Fe) are used in the buffer to increase the breakdown voltage [443], [477].
These effects are illustrated in Figure 73, showing the dispersion of the pulsed output
characteristics from a Fe-doped AlGaN/GaN HEMT.
Figure 73. (a) Pulsed ID–VD curves measured starting from different quiescent bias points, in the
OFF-state, on a sample of wafer D (highest Fe content). (b) Pulsed ID–VG curves measured
starting from different quiescent bias points, in the OFF-state [same sample as in (a)] [151].
8.7.2 Dynamic RON increase
When GaN transistors are used in power switching converters, the most detrimental trap-related
effect is the increase in the dynamic RON compared to its DC value, resulting in an undesirable
increment in power losses [346], [478].
Possible underlying mechanisms include all the electron trapping processes already described for
RF current collapse, taking also place during the off-state phase of the switch-mode operation or
during hard switching when large VDS and ID can temporarily co-exist. Channel conductivity, that
is reduced by this increase in negative trapped charge, can not be restored promptly as the device
is driven to on-state, thus leading to the dynamic RON increase.
In addition to these mechanisms, there are additional trapping processes leading to dynamic RON
increase that are specific to power transistors. These include the electron trapping into buffer traps
within the gate-drain access region induced by the substrate leakage current [346], [479]. The latter
becomes comparable with gate and source leakage currents only at the very high drain voltages
that are typically achieved in GaN power transistors only.
Carbon doping is commonly adopted in the buffer region of power transistors to increase the
blocking voltage. When this is the case, the carbon doped buffer can be the site for peculiar
trapping mechanisms, that are mainly governed by the CN acceptor state at EV+0.9 eV [124], [451],
[455]. Holes are emitted by these traps within the gate-drain access region when the device is in
the OFF state, leading to an increase in the density of negatively-ionized CN acceptors, and
therefore to a dynamic increase in the channel conductivity and RON [449], [460]. It has been
shown that both RON increasing transients under off-state bias stress and subsequent RON recovery
transients may be thermally activated with the same activation energy of 0.9 eV [480], [481]. This
has been explained as the result of a thermally activated electron capture process [346], [479] or,
alternatively, by means of a hole redistribution model [449], [482]. Recently, it was demonstrated
that discharging and charging events in carbon-doped GaN layers (GaN:C) can be governed by
transport properties of GaN:C (details can be found in [469]).
Recent works have shown that the dynamic RON can exhibit a non-monotonic dependence on the
off-state drain voltage [483]–[485], recovering to smaller values after reaching a maximum. This
behavior has been explained by means of a “leaky dielectric” buffer model: at high off-state bias,
band-to-band tunneling through the uid channel layer may promote the generation of holes, that
can be pushed to the bottom of the buffer, partially compensating the effect of the negatively-
ionized carbon acceptors [486]. A similar behavior was also attributed to holes generated by
impact-ionization, discharging the ionized CN traps and thus attenuating the dynamic RON increase
mechanism [463].
Some of the above effects are illustrated by Figure 74, showing the dynamic RON variation
measured on GaN-based power HEMTs and a schematic representation of the related processes
[487].
Figure 74. (a) Pulsed-IV curves measured at room temperature on a GaN-based power HEMT. (b)
Dependence of on-resistance on the applied trapping bias (the four lines refer to four devices sitting
on different locations on the wafer) [487]. Schematic representation of the mechanisms responsible
for non-monotonic dynamic-Ron as a function of trapping voltage
8.7.3 Threshold-voltage instabilities in isolated-gate and p-GaN transistors
To suppress the gate current, which is a requirement especially for power transistors, isolated-gate
AlGaN/GaN MIS-HEMTs have been proposed and developed [67], [488]. These include both
unrecessed or partially-recessed MIS-HEMTs, in which the gate dielectric is deposited onto the
AlGaN barrier, and fully-recessed MIS-HEMTs, where the AlGaN barrier is completely removed
under the gate region so that the gate dielectric is formed onto the GaN buffer region (while the
2DEG at the AlGaN/GaN interface survives in the two access regions).
In these transistor types, also the gate dielectric bulk and its interface with the underlying
semiconductor can act as trapping sites. The most serious effects for device performance are the
VT instabilities adding up to other trapping effects like dynamic RON. These instabilities are
typically analyzed by applying either negative gate bias (NGB) or positive gate bias (PGB) stress
voltages with zero drain-source voltage, with the aim of isolating VT effects from the drain access-
resistance ones [489].
For normally-on devices with negative threshold voltages of several volts, VT stability under NGB
is, in particular, a critical aspect that needs careful evaluation during technology development. In
normally-off devices, VT stability under PGB is instead of major concern. In the latter devices,
assessing the VT stability under NGB can be important as well, since a negative gate voltage can
be applied to switch off the transistor, in order to prevent false turn-on and ensure safe operation
against voltage spikes on the gate [490]. Moreover, NGB measurements are in any case a proxy
for the off state operation, as similar, large values of drain-gate voltage can be achieved with both
biasing conditions.
As far as NGB effects are concerned, VT shifts of either negative (i.e. VT becoming more negative
[353], [354], [491]) or positive (i.e. VT becoming more negative [472], [490], [492]) sign, as well
as of both signs depending [446], [473] on applied bias, temperature, and stress time have been
reported. The trapping mechanisms that have been held responsible for the negative VT shifts are:
(a) electron emission from interface and/or border traps [353], [354], [473]; (b) decrease in the
negatively ionized C-related acceptors in the buffer region under the gate [491], (c) hole capture
into interface traps, where holes are provided by C-related acceptors and not necessarily by a high-
field electron-hole generation mechanism [464]. Positive VT shifts have instead been attributed to:
(d) electron injection from the gate and consequent electron capture into gate dielectric traps [472],
[473]; (e) hole-induced defect generation in the gate insulator [490], [492], [493] or interface state
generation [446]; (f) Zener electron trapping into GaN traps localized under the gate edges [446];
(g) the recombination of holes provided by the C doping (and attracted to the device surface) with
electrons injected from the gate [464].
Also PGB experiments can produce VT instabilities of both signs. More commonly, positive VT
shifts are observed [354], [494], [495]. These are attributed to channel electron injection into
interface or border traps [349], [494], [495]. Negative VT shifts under PGB have instead been
explained as the result of electron removal from dielectric traps through the gate.
VT stability of devices with p-GaN gate after PGB stress is of a great concern because of the
normally-off operation [496]. By focusing on trap-related instabilities, electron injection from the
2DEG into pre-existing defects in the AlGaN barrier (also occurring in HEMT and MIS-HEMT
structures) has been reported to cause VT to shift positively. On the other hand, negative VT shifts
have been attributed to trapping of holes at the p-GaN/AlGaN interface.
Some of the above effects are illustrated by Figure 75, showing VT instabilities effects induced by
NGB (up [353]) and PGB (bottom [497]).
        Figure 75. VT instabilities effects induced by NGB (up [353]) and PGB (bottom [497]).
Within this section an overview of different techniques adopted for the characterization of trapping
phenomena will be provided. Pulsed IV characterization will initially be introduced. Even if it does
not provide a quantitative characterization of traps activation energies, it is a widely used tool to
highlight the presence of trapping phenomena. Deep-Level Transient Spectroscopy and Deep-
Level Optical Spectroscopy will then be discussed. Their combination provides the
characterization of trap levels located within the full gap of GaN. Nevertheless, specific test
structures are needed, making them difficult to apply directly on actual transistors.
Characterization techniques based on the monitoring of device drain current evolution will then be
introduced and the different methods will be commented. Monitoring of drain current for emission
process characterization supposes that the emission process takes place when the device is biased
in on-state. If traps experience said emission when the device is biased in off-state, it would be
thus impossible to properly characterize them. On-the-fly characterization overcome this issue,
allowing for the characterization of traps experiencing charge emission during off-state operation.
Interface states can be characterized by specific measurement based on capacitance-voltage (C-V)
and conductance-voltage (G-V). Finally, photoluminescence (PL) measurement technique will be
briefly described.
8.8.1 Pulsed IV
As previously introduced, trap characterization can be carried out by means of several techniques.
The most important measurement adopted to quickly identify the presence of trapping phenomena
is pulsed-IV (PIV) characterization. The concept of this measurement is rather simple. The device
is hold into a quiescent bias point (QBP), also called baseline, with its source terminal grounded.
Synchronous and short voltage pulses, typically in the 100 ns-10 us range with a duty-cycle in the
0.1 %-1 % range, are applied to the gate and drain terminals to evaluate the device I-V
characteristics. The basic idea is that traps are not able to reach their equilibrium condition during
the short duration of the pulses. Therefore, the measurement yields the I-V characteristics of the
device obtained with the traps filled at the condition set by the quiescent bias point. Since trap
filling condition strongly depend on the applied voltages, comparing pulsed I-V obtained at
different quiescent bias points quickly allows to evaluate the presence and in some way the amount
of trapping phenomena. A typical set of pulsed I-V is performed by comparing at least three or
more different quiescent bias points [426], [504], [505]: (i) the VGS=0 V,VDS=0 V QBP, which
sets also the reference of the “fresh” device conditions; (ii) gate-lag effect is then evaluated with a
QBP where VGS is held below the device threshold voltage and VDS=0 V; (iii) drain-lag effect is
finally evaluated by a QBP with the device in off-state conditions and large VDS. Obviously, many
other combinations can be considered by the three reported are the most used for device
characterization. PIV characterization quickly offers an overview on device operation and a
qualitative evaluation of the reduced device performance due to trapping phenomena. Some
information on trap spatial localization might also be obtained by comparing results from different
QBPs. If the device experiences a threshold voltage shift, trapping phenomena are likely to be
confined within the device buffer layer and/or the barrier or insulator layer under the gate terminal
[481]. On the other hand, a decrease in its transconductance might be related to an increase in
access region resistance which might be induced by surface, barrier, or buffer traps [426].
8.8.2 DLTS/DLOS
Trap investigation has been an important topic since the beginning of the semiconductor device
development. Capacitance transients are used to obtain information about an impurity level in the
depletion region of a Schottky barrier or p-n junction. The measurement consists in observing the
capacitance transient associated with the return to thermal equilibrium of the occupation of the
level, following an initial nonequilibrium condition [506], [507] [Williams_JAP_1966,
Sah_SolidState_1970]. One can thus measure the time constant of this capacitance transient as a
function of temperature and obtain the activation energy for the level. An alternative method
named deep-level transient spectroscopy (DLTS) was introduced in [508] (see Figure 76). DLTS
is still based on capacitance transients but it allows a faster characterization. The essential feature
of DLTS is the ability to set an emission rate window such that the measurement apparatus only
responds when it sees a transient with a rate within this window. Thus, if the emission rate of a
trap is varied by varying the sample temperature, the instrument will show a response peak at the
temperature where the trap emission rate is within the window. These emission rates are thermally
activated and by the principle of detailed balance can be given as:
                                𝑒1 = (σ1 𝑣1 𝑁𝐷1 /𝑔1 ) 𝑒𝑥𝑝(−Δ𝐸/𝑘𝑇)
where σ1 is the minority-carrier capture cross section, 𝑣1 is the mean thermal velocity of minority
carriers, 𝑁𝐷1 is the effective density of states in the minority-carrier band, 𝑔1 is the degeneracy of
the trap level and Δ𝐸 is the energy separation between the trap level and the minority-carrier band.
Figure 76: Majority-carrier pulse sequence which is used to produce a capacitance transient for a
majority-carrier trap. The energy-vs-distance diagrams (with band bending omitted for simplicity)
show the p+n junction depletion region (edges denoted by shaded lines) as well as the capture and
emission processes and trap occupation before, during, and after a majority-carrier pulse [508]
While DLTS is still one of the most used techniques, it also shows some limitations. Impurity
levels with a large activation energy result in a low emission rate, leading to transients with large,
time constants whose evaluation could be masked by slow drifts [509]. A novel methodology was
thus introduced in [510] with the name of deep-level optical spectroscopy (DLOS). DLOS is based
on photostimulated capacitance transients measurements after electrical, thermal, or optical
excitation of the sample. By increasing the energy of the photons impinging on the device, the
optical cross section for the transition between the deep level and the conduction and valence band
can be extracted. DLOS can provide information not only about the ionization energies of the
levels, but also about the electron-phonon interaction and temperature dependence of the levels,
i.e., about their relations with each band [510].
Due to its wide band-gap, the characterization of GaN material has typically carried out by
combining both DLTS and DLOS measurements. DLTS is adopted for evaluating levels with
activation energies below approximately 1 eV, while deeper levels are typically investigated by
means of DLOS [455], [511], [512]. The combination of the two measurement techniques allows
a full investigation of the trap level within the whole GaN band-gap.
ID-DLTS [514], [515] is typically performed by applying a constant VDS voltage, to have a
readable drain current. The gate terminal is periodically pulsed between a filling condition and a
sensing condition (see Figure 77). The sensing condition is typically slightly above device
threshold voltage, with the aim of obtaining a readable current level while at the same time
reducing device self-heating effects. Monitoring the current transient over a certain temperature
range allows to obtain a plot of the DLTS signal vs. temperature. The detection of peaks in DLTS
spectra corresponds to the presence of a trap level, whose activation energy can be extracted by
means of an Arrhenius plot [514], [515].
Figure 77: Schematic illustration of response of the depletion layers: (a)–(d)represent the response
of the depletion layer at the timing of tA1, tA2, tB1, and tB2 on the gate bias pulse, respectively [515]
8.8.3.2 CID-DLTS/DLOS
The CID-DLTS/DLOS [409], [516] technique is based on a dynamic control of either the device
gate or drain voltages with the aim of maintaining a fixed drain current in response to a thermally
and optically stimulated trap emission. For a gate-controlled method, a constant and large enough
VDS is applied to the device to have it working in its saturation region. The gate terminal is pulsed
to the on-state to induce trap filling, and then lowered near pinch-off conditions. The dynamic
control on VGS for maintaining a constant IDS will thus give rise to a VGS transient, that will thus
be recorded. This condition is sensitive for trap levels located beneath the gate contact and
affecting device VTH. For a drain-controlled method the device is biased in its linear region with
low VDS and a VGS large enough to have a negligible forward transconductance gm. Filling
condition is here obtained by biasing the device in pinch-off conditions with large VDS. The
subsequent change in device resistance ΔRD=ΔVDS/IDS is then recorded. This condition is
particularly sensitive to traps located in the device access regions. The detection of peaks in either
the temperature or optical spectra corresponds to the presence of a trap level, whose activation
energy can be extracted by means of an Arrhenius plot [409], [516].
DCTs can be recorded with a sensing condition lying in the linear [404], [520], [521] or in the
saturation region [426], [522], [523]. Significantly different results might be obtained when
comparing results obtained in the linear and saturation region [517]. Nevertheless, an accurate
evaluation of the different DCTs response might be useful for spatially localize traps as suggested
by [516]: a sensing condition in the linear region of the device with reasonably high VGS should
highlight the effect of traps located within the device access regions. On the other hand, sensing
in saturation near the pinch-off voltage, i.e. at low current level so that the effect of device access
region can be minimized, should be more sensitive to the effect of traps located beneath the gate
contact. Other phenomena that might significantly affect the results of DCTs analysis could be the
presence of mechanisms affecting the DCTs time constants, i.e. device self-heating and electric
field enhanced emission mechanisms. Self-heating by raising the local device temperatures
enhances carrier emission process leading thus to an erroneous evaluation of the emission process.
In other words, the de-trapping appears faster than it should be at the applied base-plate
temperature [524]. Similarly field enhanced emission mechanisms such as Poole-Frenkel
phenomena [525] will lead to an error like that introduced by device self-heating. Therefore, DCTs
results needs to be carefully reviewed to avoid an erroneous estimation of the activation energy
associated with the trap level causing the observed DCTs. Based on the comments and issues
reported in this section, the authors would like to suggest that, to have an accurate estimation of
the trap activation energy, VDS voltages applied during the sensing condition should be as low as
possible, in order to reduce self-heating and electric field related enhanced emission. The use of
pulsed-drain current transient (P-DCT) methodologies can also be effective to this aim, since the
device is kept at zero bias during recovery, and short voltage pulses are used to sense the on-
resistance during the de-trapping phase [518].
The extraction of time constants at different temperatures is the process leading to the construction
of the Arrhenius plot from which the trap activation energy is obtained. Several techniques can be
used to this aim. A multi-exponential approach [520] can be adopted by least-square fitting the
experimental data with the sum of several exponential functions, typically in the amount of few
hundreds, with different amplitudes and time constants as described in the following equation
                                          𝑛
                                𝐼𝑓𝑖𝑡𝑡𝑒𝑑 = ∑ 𝑎𝑖 𝑒𝑥𝑝(−𝑡/τ𝑖 ) + 𝐼∞
                                         𝑖=1
By plotting the amplitude coefficients 𝑎𝑖 vs the time constants τ𝑖 , the traps associated time
constants can be inferred by looking at the peaks of the spectra obtained, see for example Figure
79.
Figure 79: (a) Recovery transient of IDlin and (b) corresponding time-constant spectrum at−20◦C
after applying a 1 s VDS=0 and VGS=−10 V trapping pulses [520]
A different approach relies on the so called stretched exponential fitting [339], [426]. This method
least-square fits the experimental data by means of stretched exponential functions, one for each
of the low-pass or high-pass transitions observed in DCTs according to the equation:
                                             𝑛
where n here is corresponds to the number of transitions observed in DCTs which typically ranges
between 1 and 3. Note that each exponential function corresponds to potentially a different charge
emission/capture process. The Arrhenius plot of each of the i-th process is then built using the τ𝑖
values obtained at different temperatures.
An alternative approach widely used is the analysis of the derivative of the DCT by the logarithm
of time [404], [521]. Peaks in the derived spectra are used to locate the characteristic time constants
of different traps and used to derive the emission energies.
                                        2        𝐺𝑚,𝑚𝑎𝑥 /𝜔
                               𝑁𝑠𝑠 =
                                       𝑞 𝐴 𝐺𝑚,𝑚𝑎𝑥 2        C 2
                                          (       ) + (1 − m )
                                            ω 𝐶𝑜𝑥          Cox
where q is the electronic charge, A is the area of the capacitor, Gm,max is the peak value of
conductance, ω = 2πf where f is the measurement frequency, Cox is the capacitance in accumulation
region and Cm is the capacitance corresponding to Gm,max. This method has been successfully
applied to the characterization of GaN based devices [532], [533], in which the G-V measurement
[530] relies upon the assumption that energy losses and leakage currents in dielectrics are
negligibly small [534]. To this end, quasi-static capacitance–voltage (QSCV) measurement [535]
is rarely used in AlGaN/GaN HEMTs, because of the high leakage current associated with the
Schottky gate [348]. To overcome the oxide leakage problem in GaN-based structures, high
frequency C-V measurements (HFCV) are typically adopted [536]–[538]. Nevertheless, standard
HFCV measurements performed at room temperature (RT) may not be adequate to characterize
wide bandgap GaN and AlGaN interfaces [539] because of the extremely long time constants of
deep interface traps. In fact, said deep levels are frozen at RT, thus requiring elevated temperature
C-V measurement to accurately characterize the interfaces quality [532]. Recent papers pointed
out that the conductance method may show limitations on AlGaN/GaN metal-insulator-
semiconductor capacitors, and discussed the related implications [540] Other papers [541]
proposed methodologies to analyze the interface state density of dielectric/GaN MIS devices. The
wide bandgap of GaN limits hole generation at room temperature, allowing measurements in deep
depletion. By a photoassisted high-frequency capacitance-voltage characterization, it is possible
to measure the total interface state density throughout the bandgap, by using an above bandgap
light source.
8.8.6 Photoluminescence (PL)
Photoluminescence spectroscopy (PL) is a commonly used technique for studying optical
properties, measuring band-gap energy, determining phonon modes and identify energy levels due
to impurities or defects [542], [543]. During PL measurements, a light with fixed energy above the
energy-gap of the material being characterized is directed on the sample under test. Luminescence
is then detected at different wavelengths and peaks can be observed in correspondence of allowed
(radiative) transitions within the energy-gap of the semiconductor. One of the most famous result
obtained by PL on GaN layer has been the so called yellow-band YL [544], [545]. PL is a relatively
fast, contactless, and nondestructive technique and can provide very high spatial resolution. It thus
offers the possibility to evaluate material properties from samples of various sizes (from microns
to centimeters) without having to undergo a complex sample preparation [546].
9 Degradation processes in GaN devices
The possible device types and structures based on gallium nitride have a huge variety, as well as
many applications, and therefore a high number of possible degradation and failure modes may
take place. In this section, we will review the most relevant ones, with reference to lateral and
vertical devices. The discussion will be organized based on the operating conditions that lead to
the detected effects. In the first part, the ON-state and OFF-state bias conditions will be discussed,
since they are the most relevant ones for a power device. Then, the transition between the two
phases will be considered, leading to additional degradation in the          SEMI-ON-state.   Finally,
reliability in realistic environment will be analyzed, including the study of electrostatic discharge
(ESD) and electrical overstress (EOS) events, and radiation hardness.
9.1 ON-state
The on-state bias condition can be heavily stressful for several reasons. A large gate overdrive is
usually desired, in order to increase the channel carrier density and, therefore, the maximum
current levels, but this leads to a high voltage drop and electric field across the gate stack.
Moreover, although the desired voltage drop on the device should be zero in on-state, to ensure
maximum power transfer and no power loss, real devices may have voltage drops in the range of
hundreds of millivolts. At the high operating current levels, this may cause a significant power
dissipation and self-heating, that can be detrimental for long-term operation.
A high gate overdrive can be dangerous for insulated gate devices, where the electric field in the
gate dielectric has to be engineered in order not to exceed the breakdown value.[547]
Even when this design rule is correctly achieved, the insulator can still show some degradation in
its performance over time, due to a physical process called time-dependent dielectric breakdown
(TDDB) and usually linked to defect percolation.[548] As shown in Figure 80 (a), before stress
some defects are already present in the bulk of the oxide. When the device is stressed (Figure 80
(b)), additional defects can be created by the high electric field or by the current flowing through
the insulator. The spatial distribution of these defects is, in principle, random, but the presence of
a pre-existing defect can alter the local energy configuration, leading to a higher electric field and,
therefore, to a larger defect creation probability. Once enough defects are created and link up in a
complete conduction path, as shown in Figure 80 (c), a large current can flow through the oxide,
leading to a loss in isolation and to the possible catastrophic failure of the oxide if the power
dissipation is too high.
 Figure 80: sketch of time-dependent dielectric breakdown caused by defect percolation. [548]
Since the defect-creation process is field-assisted, operation of a device at a larger gate overdrive
causes a faster degradation, as commonly observed; in reliability tests a β parameter of the Weibull
distribution larger than 1 may be extrapolated, consistently with a degradation related to intrinsic
causes.[549], [550] Before catastrophic failure occurs, detected effects on the device performance
include increase in the gate leakage current due to the creation and destruction of the metastable
conduction paths (see Figure 81) [551] and the increase in gate leakage absolute value,[552] but
the effects on the device performance in terms of threshold voltage and drain current are
minor.[552] The possible presence of hole trapping in the oxide, originated by impact ionization
under the high electric field, has also been speculated.[550]
 Figure 81: Increase in gate leakage and metastable behavior caused by ON-state stress in a MIS-
 HEMT. [551]
The TDDB can affect not only lateral devices but also vertical ones, such as trench
MOSFETs[320], [553] and FinFETs.[554] The results quoted above demonstrate the possibility
of using electroluminescence measurements to pinpoint the location of the failure spots in the
device, and of limiting the current during the catastrophic failure to avoid extensive damage to the
sample, leaving a failure analysis feasible on the detected spot. Additionally, a careful design of
the insulator composition and bilayer structure can significantly improve robustness and
lifetime.[553] In the case of trench transistors, numerical simulations show electric field crowding
at the trench edges,[320] whereas in the FinFETs it is located at the corner of the dielectric, at the
foot of the fin [554] (Figure 82).
 Figure 82: numerical simulations of the electric field in (top) trench MOSFETs and (bottom)
 FinFETs. [320], [554]
In order to achieve normally-off operation, the most common approach is to use a p-GaN gate
layer, as discussed in Section 6.2.4. Under the strong gate overdrive of typical operating conditions
and with the channel formed, a large voltage drop and electric field is present in the thin p-GaN
layer, leading to possible reliability issues.
When a stress below the catastrophic failure is applied, a time-dependent degradation process
might still be present, leading to sudden jumps and a larger noise in the gate current level.[555]
The jumps are associated to the appearance of additional electroluminescence spots at the gate
edge, suggesting that conductive paths are created in the gate stack due to a defect generation and
percolation process, similar to what was found in the case of dielectrics in the previous
section.[556] Early reports suggested that the traps are created due to impact ionization of electrons
injected from the channel into the p-GaN region and accelerated by the high electric field.[557]
This idea still needs to be supported by spectral electroluminescence measurements; so far, no
band-to-band recombination and only bremsstrahlung and yellow luminescence,[556] or other
sub-bandgap wavelengths have been detected.[558]
An additional similarity with the dielectric degradation is the exponential dependence of the time-
to-failure on the stress voltage, which is also found to be Weibull-distributed. The shape parameter
β has been reported to be lower than 1, suggesting an extrinsic breakdown mechanism,[556], [559],
and greater than 1, as in recent reports.[558], [560] The degradation should happen at the edge of
the gate, as suggested by the aforementioned EL measurements and by tests of devices with
different gate width and gate length.[561] This is supported by results of numerical simulations
(Figure 83), showing that the electric field value is maximum at the edge of the p-GaN, lower in
the center of the p-GaN and furthermore reduced in the AlGaN barrier, which should therefore not
cause any reliability issues.[556]
 Figure 83: Numerical simulations of the electric field distribution in the p-GaN and AlGaN
 barrier during high gate bias stress. [556]
Even though the degradation is found to be faster at higher temperature,[561] a different behavior
can be seen in case of more complex degradation processes.[562], [563] In [562], the degradation
is supposed to be enhanced by the accumulation of positive charges at the p-GaN/AlGaN interface,
which promotes the injection of electrons from the channel into the p-GaN, with consequent
degradation. Higher temperature leads to more hole release, therefore improving the reliability.
The presence of the hole trapping is confirmed by the fact that the gate current decreases during
stress, due to the electrostatic repulsion between the injected and trapped holes. This degradation
process is sketched in Figure 84, which also shows the electroluminescence distribution and the
correlation between the electroluminescence intensity and the gate current. It is worth mentioning
that carrier acceleration and impact ionization are also less prominent at high temperatures, due to
the increased scattering, and this can also contribute to the positive temperature coefficient.
 Figure 84: (a) spatial distribution of the electroluminescence and (b) its correlation with the gate
 current. (c-d) sketch of the degradation process leading to a higher robustness at high
 temperature in p-GaN gate devices. [562]
In the case of vertical p-n--n diodes and vertical junction field-effect transistors, most of the failures
in common reliability tests (HTRB, HTOL, THB, TC) are related to the substrate quality, the
substrate miscut angle and the morphology of the surface after the epitaxial growth.[564]
More detailed studies on pn vertical diodes show that forward bias stress induces an increase in
series resistance, in turn-on voltage and in forward and reverse leakage current, as well as a
reduction in the optical power emitted by diode due to band-to-band recombination in forward
bias.[565] The variation in series resistance and optical power are well correlated, and their
variation has a square root dependence on time, as shown in Figure 85. All the experimental results
can be explained according to the following model. During stress, the flow of carriers, aided by
the temperature, can break the residual Mg-H bonds that are still present after the Mg activation
phase done during growth. Hydrogen interstitials can then diffuse following the concentration
gradient towards the n-type material. The diffusing hydrogen can passivate Mg atoms closer to the
p-n interface, decreasing the hole concentration. This causes an increase in the turn-on voltage and
the reduction in the amount of holes available for radiative recombination. The dependence on the
square root of stress time is a common signature of diffusion-related degradation, being originated
from the solution of Fick’s second law of diffusion in one dimension.[566]
 Figure 85: (top) Schematic cross-section of the vertical GaN pn diode; (bottom) Dependence of
 the increase in the series resistance and of the decrease in electroluminescence intensity on the
 square root of stress time. [565]
9.1.4 RF stress
Operation in radiofrequency (RF) conditions is not relevant for power devices but only for
amplifiers, but a useful conclusion can be drawn by the analysis of the difference in degradation
between DC and RF conditions.
In the case of RF devices, a common degradation cause is the presence of hot electrons in near
pinch-off conditions, especially when short channel effects and poor carrier confinement play a
role due to the reduced gate length. Since the process is influenced by the source-drain leakage, a
correct compensation of the buffer conductivity can significantly improve the reliability of the
devices, with Fe and C co-doping leading to the best results.[567] The degradation mechanism is
field-driven, as suggested by tests on devices with and without field-plates.[568] A RF stress
usually induces small variations in the DC performance of state-of-the art devices, mainly a
lowering of the saturation current and a decrease in gate leakage,[569] but the generation of defects
causes an increase in the dynamic current collapse.[570] A negative[571] or positive[572] shift in
threshold voltage is also observed sometimes, due to the presence of traps in the cap and barrier
layer.[571]
The important conclusion comes from the comparison between devices stressed in DC and RF
conditions, showing different degradation modes and a stronger degradation in the RF case.[571]
This suggests that, even for power devices, it may be important to test the degradation not only in
ON-state, OFF-state   and SEMI-ON-state separately, but to test it also in the real switching pattern, to
take into account the interplay between the different mechanisms and time dependence occurring
in the various bias points of the switching locus.[573]
9.2 OFF-state
The   OFF-state   bias condition can be highly stressful for a power device, due to the high electric
field present, on average, in the blocking region and peaking in critical device regions, such as the
sharp corners and the gate edge. The electric field can cause damage to materials that are part of
the device without being the semiconductor itself, such as the insulators for passivation and gate
isolation. Other processes may involve the breakdown of gallium nitride itself. In addition,
interaction with the external atmosphere may promote field-assisted chemical reactions. In some
cases, the high electric field can also be present together with a high current density, such as during
avalanche conduction. In the following, we will discuss all these possibilities.
As discussed in section 9.1.1, degradation of the dielectrics is usually present in      ON-state   stress
conditions, since the voltage drop across the gate insulator is the highest, whereas in        OFF-state
stress the large voltage drop in the depleted semiconductor mitigates the electric field across the
gate stack. Therefore, some of the results already presented in section 9.1.1 still hold, even when
a larger bias is applied: the better reliability of engineered bilayer insulators,[320], [553] the
location of weak spots corresponding to electric field peaks[554], and the negative[547] or
positive[574] threshold voltage shift related to defect generation, which may also cause worse
dynamic performance.[547]
Nevertheless, additional dielectrics are present in the complex structure of a state-of-the-art GaN
power devices, such as the ones used for passivation and isolation of the field-plates. Their
degradation results in an increase in gate leakage level and noise, eventually leading to catastrophic
failure even significantly below the breakdown voltage of the device; dielectric degradation is
typically characterized by an exponential dependence of the average time to failure on the stress
voltage.[575] In this case, the shape parameter β of the Weibull distribution is lower than 1,
suggesting that reliability is limited by extrinsic factors. A possible origin for this experimental
behavior is the degradation of the silicon nitride used for the passivation and field plate isolation.
As can be seen through numerical simulations (Figure 86 (a)), in the case described in [575], in
off-state conditions the peak of the electric field is located in the silicon nitride, in proximity of
the edge of the gate overhang on the drain side, and the value is ~6 MV/cm, comparable with the
theoretical breakdown electric field (∼ 6 MV/cm). [575] The electric field in the AlGaN barrier is
lower than 3 MV/cm, therefore degradation in this layer is unlikely. This hypothesis can be
confirmed by locating the failure spot by means of electroluminescence measurements (Figure 86
(c)) and performing cross-sectional transmission electron microscope (TEM) failure analysis on
it.[139] As can be seen in Figure 86 (b), the damaged region starts at the edge of the gate region
and extends into the channel region.
   (a)
(b)
(c)
 Figure 86: (a) TEM cross-section analysis is able to identify the degradation caused by (b) the
 peak in electric field at the edge of the gate head (as obtained by numerical simulations),
 located by means of (c) electroluminescence measurements. [575], [139]
The mechanism can be further confirmed and solved by changing the vertical conductivity in the
epilayer, leading to a different grounding of the floating buffer and, therefore, to a larger 2DEG
retraction in the case of the improved devices, as shown in the numerical simulations in Figure 87
(d).[575] This leads to a significantly lower electric field in the silicon nitride for the second
generation of devices (Figure 87 (a-c)) and to an improvement in the time-to-failure of more than
three orders of magnitude, as shown in Figure 87 (e).
 Figure 87: (a-c) improved devices with different grounding of the floating buffer have a lower
 electric field due to (d) the larger 2DEG retraction. (e) this change leads to an increase in time-
 to-failure by more than three orders of magnitude. [575]
A second possible solution is to change the length of the gate head at the drain side, limiting the
peak electric field, or to add an extra SiN layer under the gate head.[139][576]
As discussed in Section 4, gallium nitride is a polar material. Any polar material, when submitted
to a compressive or tensile mechanical stress, builds a potential across itself, and this behavior is
called direct piezoelectric effect. In polar materials the opposite is also true. An external voltage
applied in the direction of the polarization vector may cause an expansion or a contraction of the
material: this is called the converse (or inverse) piezoelectric effect. A sketch of all the processes
is shown in Figure 88.
 Figure 88: sketch of the direct and converse piezoelectric effect in polar materials.
 http://resources.edb.hkedcity.net/physics/articleIE/smartmaterials/SmartMaterials_e.htm
This behavior is not only theoretical, but can be experimentally measured by means of
interferometric techniques on GaN crystals.[577] The amount of displacement caused by different
voltages applied to a 2.5 µm thick single crystal [0001] GaN film is shown in Figure 89, and
corresponds to a displacement of 0.05 nm at an electric field of 64 000 V/cm, i.e. to a displacement
of 1.56 nm at an electric field of 2 MV/cm. The electric field peak found close to the gate edge at
the drain side in common HEMT devices is close or can even exceed this value, even in presence
of multiple field plates, in common high power operation. Therefore, one may wonder if this level
of displacement (without considering the different value of the piezoelectric coefficients between
the reported GaN film and the AlGaN of the barrier) can cause damage to the device and limit its
reliability. Of course, a localized electric field would not be enough to produce a displacement,
since the lattice would build up strain to accommodate for the expected local displacement, but the
levels of piezoelectric stress originated by the expected variation can be very high, in the order of
hundreds of MPa.[578], [579]
 Figure 89: GaN displacement measured at various applied voltages. [577]
This idea was proposed for the first time in 2006,[580] and a lot of work has been done to identify
this mechanism in the following years. Specific tests have found the presence of a critical gate or
drain voltage for the degradation, consistent with the build-up of a critical amount of strain in the
structure before relaxation and formation of defects takes place.[580] When stressed above the
critical voltage, an increase in charge trapping inside the AlGaN barrier takes place, consistently
with the generation of defects in that region, whereas the trapping in the buffer remains
unaffected.[581] By means of stresses at different gate voltages and current levels it was possible
to exclude any effect of the hot electrons in the process, since the stress current was found to have
no impact on the critical voltage value.[582] The generation of cracks only at the drain side of the
gate taking place above the critical voltage, without any impact in the degradation of temperature
or hot electrons, was reported by TEM measurements.[583]
The generation of indentations, pits and grooves at the drain side of the gate edge was confirmed
also by atomic force microscope (AFM) and scanning electron microscope (SEM) testing of
stressed devices after chemical etching of the gate metal, as shown in Figure 90. At least some of
the new pits form without the presence of any pre-existing defects, consistently with the inverse
piezoelectric effect. The defects act as gate leakage paths, may reduce the maximum drain current
and favor the charge trapping in the device.
 Figure 90: formation of grooves and pits in devices stressed at increasing drain-gate voltage.
 [584]
When devices are stressed in reverse gate bias conditions, particles and stringers appear along the
gate edge, both at the source and the drain side, as shown in Figure 91 (a-c).[585] Their chemical
composition can be determined by Auger electron spectroscopy (AES), reported in Figure 91 (d),
highlighting an oxygen-rich composition and the presence of gallium. The quantitative analysis is
compatible with the formation of Ga2O3 and Al2O3. After metal removal, the presence of large pits
below each oxide particle is clearly visible (Figure 91 (e)). By changing the polarity of the stress
bias on the gate-source and gate-drain diodes, it is possible to demonstrate that the oxide forms
where the electric field is the highest. By testing devices in vacuum it was possible to demonstrate
that the oxide only forms when oxygen is present in the atmosphere, and it is supposed that oxygen
diffuses through the passivation layer and locally oxidizes the AlGaN/GaN layers.
 Figure 91: (a-c) top view SEM images of an HEMT stressed at reverse gate bias for increasing
 periods of time. The chemical composition of the stringers is shown in (d), and (e) shows the
 pits that form below them. [585]
Similar effects can be produced by reaction with hydroxyl groups (OH-) or water.[586] In this
case, the degradation happens due to anodic oxidation of the AlGaN layer in the electrochemical
cell composed of the gate metal, the passivation layer and the barrier, aided by water and by the
presence of holes. The main effects are a degradation in drain current and dynamic performance,
whereas in this case the increase in gate leakage is not caused by the presence of the pits.
It is also suggested that an exclusion zone can appear around the pits, where no additional
degradation spots can be formed, as the result of the consumption of mobile species in the
electrochemical reaction.[587]
9.2.2.3 Time-dependent breakdown of GaN epitaxial stacks
The results presented in section 9.1.2 for the p-GaN degradation show trends comparable to the
ones reported in section 9.1.1 for the time-dependent dielectric breakdown degradation, namely
the Weibull distribution of the failure time and its exponential dependence on the stress voltage.
One can then wonder if gallium nitride itself, when deeply depleted, can behave as a dielectric and
experience a TDDB-like degradation process. In general, detecting a TDDB process in a full GaN
power device can be complex, due to the large number of parts and materials it is composed of,
each of them possibly contributing to the degradation as discussed in this Section 9. Moreover, if
a device is not fully vertical, both a lateral and a vertical electric field are present, further
complicating the analysis. Therefore, even though the presence of GaN TDDB was suggested in
the past, no demonstration excluding other possible causes was provided.[588], [589] For this
reason, the first conclusive demonstration of time-dependent dielectric breakdown in the gallium
nitride itself was done on optoelectronic devices.[590] LEDs have an intrinsic vertical structure,
no complex parts such as field plates, gate with various geometries, blocking dielectrics or
engineered passivation, and are already at a high technology readiness level, allowing for testing
of stable and reliable devices with no reliability and behavioral issues, which is not the case for
fully-vertical power devices. The experimental tests, carried out in reverse bias condition, confirm
that gallium nitride behaves as a leaky dielectric when deeply depleted, that the time-to-failure has
an exponential dependence on the stress electric field, and that it is Weibull distributed.[590] The
extrapolated shape parameter β is 4.43, confirming that the devices under test show an intrinsic
failure behavior. Consistent results were obtained also on GaN-on-Si stacks used for the
fabrication of GaN transistors. It was demonstrated that when submitted to drain-to-substrate (2
terminal) stress, AlGaN/GaN transistors can show a time-dependent degradation, which is Weibull
distributed, significantly field-dependent and weakly thermally activated (Ea=0.25 eV).
Degradation was ascribed to a percolation process, leading to the generation of localized shunt
paths between drain and substrate. [591]
9.2.2.4 Buffer decomposition experiments
When a lateral power device is biased in the off-state, a large voltage drop is present in all the
layers from drain to substrate, which usually include the AlN nucleation layer, an AlGaN buffer
and a carbon-doped GaN buffer. Understanding which of these layers has the strongest impact on
the vertical breakdown voltage and on device reliability is a critical task for improving the final
behavior. This can be done by testing separate structures, where the growth has been stopped after
each of the single layers above the silicon substrate, to form AlN/Si, AlGaN/AlN/Si, and
C:GaN/AlGaN/AlN/Si stacks.[592] [593] This way, the variation in reliability (and charge
trapping) induced by each layer can be analyzed independently. In order to do this, a mandatory
technical requirement is the processing of good ohmic contacts on top of each layer, a non-trivial
task on low-conductivity and wide-bandgap materials.
The analysis of AlN layers directly grown on a silicon substrate indicated a breakdown field of
3.25 MV/cm, which is significantly lower than that of high-quality crystalline AlN. The difference
is ascribed to the presence (and high density) of vertical leakage paths, involving V-pits and
threading dislocations. This value further decreases with temperature (see Figure 92) [593]. To
explain this temperature dependence, the failure voltage was plotted against the leakage current.
Samples with higher leakage showed a lower breakdown voltage, and a lower breakdown field, of
the AlN nucleation layers. The flow of current at localized defects may lead to premature
breakdown. This argument was used to explain why the breakdown field of AlN grown on silicon
(3.25 MV/cm) is much lower than that of high quality crystalline AlN (up to 12 MV/cm).
In addition, it was demonstrated that the AlN/Si structures show trapping of negative charge, which
has been ascribed to the injection of electrons from the silicon substrate towards deep traps located
in the AlN. By adding an AlGaN layer on top of the AlN, it is possible to significantly reduce the
density of defects, and this results in a more uniform sample-to-sample leakage current. A strong
increase in the breakdown voltage can be reached by adding a C:GaN layer on top of the stack. In
this case, the structures analyzed in [593] showed breakdown voltages larger than 800 V. It is
worth noticing that the presence of a C:GaN layer can result in positive charge trapping. This is
ascribed to the presence of holes from C:GaN, that are trapped at the GaN/AlGaN interface,
leading to the storage of positive charge in the buffer.
In the case of the AlN/Si stacks under analysis, the leakage current originates, as confirmed by
numerical simulations, from the series connection of equivalent AlN/n+-Si and n+-Si/p-Si
junctions, due to the presence of an inversion layer at the interface.[594] The limiting factors are
the p-doping level of the silicon substrate and the electric field at the interface, since both of them
control the tunneling injection from the silicon substrate into the AlN layer.
 Figure 92: Dependence on temperature of the failure voltage of AlN/Si layers, and its correlation
 with the leakage level. [593]
All the processes discussed above can affect also vertical devices, but a degradation mechanism
exclusive to them exists. Vertical power devices are built in order to withstand high voltage levels,
as discussed in section 7, and the high electric field and long drift region makes them capable of
avalanche current conduction. This is not an unwanted side-effect but a desired feature, especially
in power diodes, since it allows them to withstand surge events.
However, one needs to consider that if the structure is not optimized, operation in avalanche mode
may be detrimental for the reliability of the devices, due to the high electric field and to the
presence of a large number of highly-energetic carriers. The former can cause failure due to non-
optimal termination and electric field peak at the bevel surface,[595] whereas the latter can create
significant lattice damage.[596]
The actual defect species responding during avalanche operation is not yet clear, but a possible
hypothesis is that carbon in nitrogen substitutional position (CN) may be involved. These defects
create a wide range of effects, including an increase in series resistance and leakage current, a
variation in the turn-on voltage, an increase in avalanche breakdown voltage (avalanche walkout
[597]) and a worsening of the dynamic performance. The analysis is made more complex by the
large variations in vertical electric field in the lateral direction, as summarized in Figure 93. In the
high-field region, the avalanche condition and the high current conduction causes damage creation
and trapping of a part of the flowing charges, whereas in the medium field region no avalanche is
present, only a negligible amount of current is flowing and therefore the main effect is field-
assisted de-trapping of native charge. When the device is moved to measure bias condition, the
trapped carriers cause the increase in avalanche breakdown voltage, the de-trapped ones modify
the turn-on voltage, and the lattice damage impacts on the leakage current and the series
resistance.[596]
 Figure 93: summary of the possible effects taking place in a vertical GaN power diode during
 stress in avalanche conduction. [209]
9.3 SEMI-ON-state
In real operating conditions, a power switch is not limited only to      ON-state   and   OFF-state   bias
points. During the turn-ON and turn-OFF transitions, the presence of the internal and parasitic
capacitances prevents an instantaneous variation of the voltage at the terminals of the device,
therefore a relatively high voltage at the drain can still be present when the gate voltage is already
close to the threshold value. This bias condition is called   SEMI-ON-state,   because the conductive
channel is partially formed. Depending on the residual drain voltage during the transition, in SEMI-
ON-state   state a relatively large amount of electrons in the channel can be accelerated by a large
electric field, causing the flow of highly-energetic (hot) carriers in the device that can be
detrimental for its reliability.
The main effects a     SEMI-ON-state   stress can cause on a transistor include a worsening of the
isolation properties of the gate diode, threshold voltage shifts, increase in ON-resistance, decrease
in transconductance peak value, change in peak electric field, knee walkout and increase in drain-
lag transient amplitude [598], [599]. These changes may be ascribed to the increase in defect
concentration close to the channel region, shown in Figure 94, caused by the energy exchange
between the hot electrons and the crystal lattice, as can be detected by several deep level
characterization techniques applied during stress [599]. The specific microscopic configuration of
the deep level causing the detected variations can change depending on the device growth
conditions, structure and processing. Recent papers investigated the effect of hot electrons on the
dynamic on-resistance of AlGaN/GaN high-electron mobility transistors subject to semi-on stress.
The additional dynamic-Ron detected in semi-on state was ascribed to hot-electron trapping at the
passivation/AlGaN interface [600].
Specific circuits [434], [601]–[603] can be used to test the degradation of GaN HEMTs induced
by hot-electrons. A promising approach was proposed in [603], [604], based on the use of
switching setups capable of monitoring on-wafer the effects of semi-on stress. By tuning the
capacitance at the drain node, it is possible to control the amount of energy/charge released during
hard switching events, and to evaluate the effect of stress on the devices. The comparison of results
obtained via different techniques [518], indicated that hot-electron trapping is a very fast process,
that takes place in few nanoseconds. The developed methodologies allowed to assess the trajectory
of the switching locus, the power dissipated during the turn-on transitions, and the dynamic on-
resistance of the devices (see an example in Figure 95). The kinetics of hot-electron trapping were
modeled analytically in a recent publication, based on rate equations [605].
 Figure 94: increase in amplitude of the drain-lag transient caused by        SEMI-ON-stress,   and
 corresponding increase in defect density detected by CID-DLOS. [409]
Figure 95: (left) Switching locus curve (a) comparing purely resistive switching and hard switching
(b) applying different stress voltages (VDD). The trajectory confirms the hard switching condition
during the turn-on transition. (right) Power dissipated during the turn-on transition. The power
peak increases with increasing the stress voltage VDD [518](PERMISSION FROM
https://iopscience.iop.org/article/10.1088/1361-6641/abc456/pdf)
In the specific case of gate injection transistors (GITs), a destructive positive feedback loop
initiated by the hot electrons created during   SEMI-ON-stress   can lead to a rapid increase in the
electric field at the drain side and to the failure of the device. [573][606] This issue can be solved
by using hybrid drain gate injection transistors (HD-GITs) device configuration, since the hole
injection from the p-GaN at the drain partially reduces the amount of trapped charge caused by the
hot electrons and the electric field peak.[573]
Dangerous bias points for the reliability of a device can originate not only from the expected
behavior during operation, but also from external unwanted stimuli. A frequent example is the
application of short voltage or current pulses to one of the terminals, nanoseconds to seconds long,
as a consequence of the interaction with users, the ambient conditions, or machinery during
fabrication, assembly and handling. Collectively, these phenomena are called electrostatic
discharges (ESD) and electrical overstress (EOS).
Under such high power dissipation conditions, several parts of the device can show degradation.
One of them is the ohmic contacts at the source and the drain, as can be evaluated by tracking their
specific resistance without any contribution by the channel resistance [607]. At increasing current
stress levels in floating gate condition, the values of the contact resistance can significantly
increase, up to a factor of ≈ 30. Backside infrared camera measurements, reported in Figure 96,
show the creation of dark spots, attributed to electromigration leading to current filamentation. The
spots are always present on the grounded contact, suggesting that not only the high power
dissipation but also the current flow or electric field direction plays a role in the degradation. In
this case, no variation in material parameters, charge density or electron mobility is observed.
 Figure 96: backside infrared camera measurements, showing the appearance of dark spots on
 the contact (source in part a, drain in part b) grounded during the ESD event. [607]
If the gate reliability is not the main goal of the analysis, it is possible to test devices without the
gate metal [609]. If a mesa-type isolation is not present, the migration of the contact metal from
drain to source, leading to a drain-source short, is a relevant process, especially at the corners of
the contacts, where the electric field is the highest. The process is started by the melting of the
metal at the high power dissipation condition, which is expected to happen at temperatures 1.5x to
4x lower than the GaN one. If a mesa isolation is present, the short-circuit instead forms along the
channel, due to the presence of surface states that act as low-energy paths for the metal migration.
One of the main shortcomings with ESD/EOS testing on transistors is the fact that a single pulse
is applied to the device under test, whereas it was demonstrated that GaN-based devices suffer
from cumulative effects [609]. Another problem is that the protocol is usually based only on gate-
floating and gate-grounded configurations. Unfortunately, in the real application the gate voltage
will sweep the entire operating range, and there is no way to guarantee that the ESD event will
happen only when the device is biased in one of the two tested configurations. For this reason, in
a three-terminal device it is important to analyze the dependence of the robustness on the applied
gate voltage [610].
As shown in Figure 97 (a), when the ESD event is applied in          OFF-state   and   SEMI-ON-state,   the
catastrophic failure of the device always takes place at the same voltage on the device under test.
This finding is consistent with a failure related to reaching the breakdown electric field of the
material, which always happens at the same drain-source voltage regardless of the applied gate
voltage. When the ESD event is applied in       ON-state,   the detected robustness is lower, and the
failure voltage and current points closely follow an iso-power curve. This behavior confirms that,
in the on-state, the reliability is limited by the power dissipation that the device can withstand.
These assumptions are confirmed by the results of a scanning electron microscope (SEM)
inspection (Figure 97 (c), for a device failed after testing in ON-state, and Figure 97 (d), for a device
tested in OFF-state). The energy-dispersive X-ray spectroscopy (EDX) analysis in correspondence
of the failure points (red crosses in Figure 97 (c-d)) is reported in Figure 97 (b). The failed region
of the device tested in ON-state has a high gallium and aluminum content, possibly originated by
the melting of the barrier and channel layers. The crack in the device tested in OFF-state has a lower
gallium ad aluminum concentration, since the GaN-based layers are separated side-by-side, and a
high silicon signal, because the electron beam is able to reach a larger part of the silicon substrate
through the crack [610], [611].
 Figure 97: (a) failure caused by critical electric field and power dissipation in a device submitted
 to drain ESD discharges at various gate voltages. EDX measurements in (b) of the failed region
 in a device failed due to (c) power dissipation and (d) electric field confirm the driving force of
 the failure. [610], [611]
Owing to their very high operating limits and reliability, GaN-based devices are attractive for a
wide range of special uses in extreme ambient conditions, such as space applications. In order to
be a viable option, their robustness against radiation doses has to be analyzed and confirmed.
Different types of irradiation can lead to different effects, therefore in the following we will
analyze each of them separately. For an interested reader, an extensive discussion can be found in
recent reviews [612]–[614].
The irradiation with protons can cause a wide range of effects in GaN-based devices, which can
be related to the creation of new defects (or to the increase in concentration of pre-existing ones)
caused by the interaction between the crystal lattice and the energetic protons.
The increase in defect concentration leads to a lowering of the mobility of the material, due to the
scattering between the carriers and the defects, and to a reduction in carrier density, caused by the
charge trapping in the defects and by the corresponding electrostatic effects. On the final devices,
this corresponds to an increase in on-resistance, a positive shift in the threshold voltage, a reduction
in the saturation current and in the peak transconductance value. The larger defect concentration
causes a worsening of the dynamic behavior, due to increased density of states available for the
trapping of charge.
In some cases, the dynamic RON is found to be decreased rather than increased by the proton
irradiation [615]: on irradiated devices an increase in off-state leakage is observed, indicating an
increase in the unintentionally-doped GaN layer conductivity. This conductivity increase leads to
an increased deionization rate, thus reducing the dynamic on-resistance.[615]
The level of damage depends on the energy of the protons, and it is found to be higher for lower
energy [616], [617]. This behavior may seem unexpected, but is related to the larger non-ionizing
energy loss of the lower energy particles in the barrier and channel region, whereas higher energy
protons cause damage deeper in the device structure. Concerning the number of irradiated protons,
a typical threshold value (at 1.8 MeV) for the variation in device performance is a fluence of 1012-
1013 p+/cm2 at 1.8 MeV[618]–[620], 5×109 p+/cm2 at 40 MeV[621]. Based on the energy
dependence discussed above, the lower threshold fluence at higher energy is unexpected, and may
be related to different device structure and quality in the various papers. It is worth pointing out
that values for variation in dynamic performance may be lower than the ones reported for variation
in DC characteristics[622].
Specific defects created by proton irradiation include both deep donor and deep acceptor levels, as
well as de-hydrogenation and reconfiguration of ON-H complexes.[623], [624]
Most of the degradation caused by proton irradiation is related to displacement damage and is
recoverable by high temperature annealing.[625], [626]
Neutron irradiation produces two main effects: an increase in threshold voltage and a decrease in
the material mobility. The threshold voltage variation originates from the increase in concentration
of pre-existing acceptor states in the barrier, and the amount of variation increases linearly with
the neutron fluence. [612]–[614] The cause of the lower material mobility is attributed to barrier
acceptors too, which cause local threshold voltage and channel density variations and, therefore,
additional scattering.
In general, neutrons are found to be less damaging than protons, both in terms of         ON-resistance
and breakdown voltage,[623] but cause extended damage rather than the creation of point defects
found by other irradiation species.[612]
The main effect of electron irradiation is to cause a negative threshold voltage shift at lower
fluence, which becomes positive at higher fluence, originated by the generation of traps with
different charge states. Additionally, a monotonic decrease in channel mobility has been reported.
The defects responsible for these variations have been extensively studied, and include nitrogen
vacancies and additional traps with 0.3 eV, 0.45 eV, 0.55 eV and 0.8 eV activation energy.[613]
The irradiation with gamma rays is less tested, but some information is available from the
literature. Reported effects include increase in   ON-resistance,   negative shift in threshold voltage,
increased saturation current, decreased gate current and increased reverse breakdown
voltage[627]. The generated defects are supposed to be nitrogen vacancies (VN), which act as
donors in gallium nitride. Conflicting results have been reported, indicating instead a reduction in
current[628], [629]. The difference may originate from different device structure and quality, or
from the impact of different dose levels, showing improvements at lower dose and damage at high
dose.[630]
Additional ionizing species have been tested on gallium nitride devices, including helium and
carbon.[631] Effects include increase in   ON-resistance,    positive shift in threshold voltage and
reduction in saturation current, linearly proportional to the fluence, with no influence of the particle
mass or energy. No variation in gate leakage was found.
10 Conclusions
In summary, with this tutorial paper we reviewed the properties and advantages of gallium nitride,
and the characteristics, technology and reliability of GaN-based transistors. These devices are
expected to play a significant role in next-generation power converters: competition with silicon
and silicon carbide is expected to become stronger and stronger in the coming years. A substantial
improvement in device technology will be possible through extensive research, both at academic
and industrial levels.
If on one hand current GaN transistors can significantly improve the performance and reliability
of switching mode power converters, next-generation scaled devices will enable monolithic
integration, thus paving the way for the fabrication of miniaturized and MHz-range power
converters. Further competition with silicon and SiC will come from the development of vertical
GaN transistors, which have a great potential for high power/high voltage applications, and can be
fabricated on inexpensive and large substrates.
A deep understanding of material and device properties will allow to further extend the operating
ranges of the devices, thus enabling robust kV-range operation (for large-size devices) and sub-
microsecond switching (for scaled transistors). Finally, a deep knowledge of trapping and
reliability-limiting processes will allow to push the devices to the limits, thus fully exploiting the
competitive advantage of GaN, as a ground-breaking semiconductor for power electronics.
This paper provides a comprehensive overview on the properties of GaN and related devices, and
can be used as a reference for researchers willing to enter this interesting and complex field, or
looking for an update on the most recent advancements on device technology.
11 Acknowledgments
This project has received funding from the ECSEL Joint Undertaking (JU) under grant agreement
No 826392. The JU receives support from the European Union’s Horizon 2020 research and
innovation programme and Austria, Belgium, Germany, Italy, Slovakia, Spain, Sweden, Norway,
Switzerland.
This project has received funding from the ECSEL Joint Undertaking (JU) under grant agreement
No 876659. The JU receives support from the European Union’s Horizon 2020 research and
innovation programme and Germany, Austria, Slovakia, Sweden, Finland, Belgium, Italy, Spain,
Netherlands, Slovenia, Greece, France, Turkey. In addition, Germany including the Free States of
Saxony and Thuringia, Austria, Belgium, Finland, France, Italy, the Netherlands, Slovakia, Spain,
Sweden, and Turkey provide national funding.
The project leading to this application has received funding from the ECSEL Joint Undertaking
(JU) undergrant agreement No 101007229. The JU receives support from the European Union’s
Horizon 2020 researchand innovation programme and Germany, France, Belgium, Austria,
Sweden, Spain, Italy
12 Reference
[1]   T. Morita et al., “99.3% Efficiency of three-phase inverter for motor drive using GaN-based
      Gate Injection Transistors,” in 2011 Twenty-Sixth Annual IEEE Applied Power Electronics
      Conference and Exposition (APEC), 2011, 481–484, DOI: 10.1109/APEC.2011.5744640.
[2]   D. Han et al., “Efficiency characterization and thermal study of GaN based 1 kW inverter,”
      in 2014 IEEE Applied Power Electronics Conference and Exposition - APEC 2014, 2014,
      2344–2350, DOI: 10.1109/APEC.2014.6803631.
[3]   W. Zhang et al., “Evaluation and comparison of silicon and gallium nitride power transistors
      in LLC resonant converter,” in 2012 IEEE Energy Conversion Congress and Exposition
      (ECCE), 2012, 1362–1366, DOI: 10.1109/ECCE.2012.6342657.
[4]   E. Gurpinar et al., “Ultra-low inductance design for a GaN HEMT based 3L-ANPC
      inverter,” in 2016 IEEE Energy Conversion Congress and Exposition (ECCE), 2016, 1–8,
      DOI: 10.1109/ECCE.2016.7855540.
[5]   E. Gurpinar et al., “600 V normally-off p-gate GaN HEMT based 3-level inverter,” in 2017
      IEEE 3rd International Future Energy Electronics Conference and ECCE Asia (IFEEC
      2017 - ECCE Asia), 2017, 621–626, DOI: 10.1109/IFEEC.2017.7992110.
[6]   M.-X. He et al., “A 1-MHz GaN Converter with 4X Voltage Range *,” in 2019 IEEE
      Applied Power Electronics Conference and Exposition (APEC), 2019, 2349–2355, DOI:
      10.1109/APEC.2019.8722186.
[8]   E. Aklimi et al., IEEE J. Solid-State Circuits, 52, 6, 1618–1627, 2017, DOI:
      10.1109/JSSC.2017.2672986.
[9]   A. Villarruel-Parra et al., “75 MHz discrete GaN based multi-level buck converter for
       envelope tracking applications,” in 2019 IEEE Applied Power Electronics Conference and
       Exposition (APEC), 2019, 1553–1560, DOI: 10.1109/APEC.2019.8721823.
[10]   X. Li et al., “GaN-on-SOI: Monolithically Integrated All-GaN ICs for Power Conversion,”
       in 2019 IEEE International Electron Devices Meeting (IEDM), 2019, 4.4.1-4.4.4, DOI:
       10.1109/IEDM19573.2019.8993572.
[11]   Y. Yamashita et al., “Monolithically Integrated E-mode GaN-on-SOI Gate Driver with
       Power GaN-HEMT for MHz-Switching,” in 2018 IEEE 6th Workshop on Wide Bandgap
       Power     Devices      and        Applications          (WiPDA),         2018,       231–236,        DOI:
       10.1109/WiPDA.2018.8569057.
[12]   X. Ding et al., CES Trans. Electr. Mach. Syst., 3, 1, 54–64, 2019, DOI:
       10.30941/CESTEMS.2019.00008.
[13]   W.   S.   Tan   et   al.,   Solid.        State.   Electron.,   50,     3,     511–513,      2006,   DOI:
       10.1016/j.sse.2006.02.008.
[14]   R. Cuerdo et al., Phys. Status Solidi (c), 5, 6, 1971–1973, 2008, DOI:
       10.1002/pssc.200778555.
[15]   D. Maier et al., IEEE Trans. Device Mater. Reliab., 10, 4, 427–436, 2010, DOI:
       10.1109/TDMR.2010.2072507.
[18]   “EPC2108 - Enhancement Mode GaN Power Transistor Half Bridge with Integrated
       Synchronous           Bootstrap.”                  [Online].           Available:             https://epc-
       co.com/epc/Products/eGaNFETsandICs/EPC2108.aspx.
[24]   I. Hwang et al., “1.6kV, 2.9 mΩ cm2 normally-off p-GaN HEMT device,” in 2012 24th
       International Symposium on Power Semiconductor Devices and ICs, 2012, 41–44, DOI:
       10.1109/ISPSD.2012.6229018.
[25]   E.   Dogmus     et   al.,   Appl.   Phys.   Express,   11,    3,   034102,     2018,   DOI:
       10.7567/APEX.11.034102.
[26]   N. Herbecq et al., Phys. Status Solidi (a), 213, 4, 873–877, 2016, DOI:
       10.1002/pssa.201532572.
[27] S. Q. Zhou et al., Appl. Phys. Lett., 86, 8, 081912, 2005, DOI: 10.1063/1.1868870.
[28] A. Dadgar et al., New J. Phys., 9, 10, 389–389, 2007, DOI: 10.1088/1367-2630/9/10/389.
[30] Y. Cai et al., Materials (Basel)., 11, 10, 1968, 2018, DOI: 10.3390/ma11101968.
[32]   S. Keller et al., Semicond. Sci. Technol., 29, 11, 113001, 2014, DOI: 10.1088/0268-
       1242/29/11/113001.
[33]   J. L. Hudgins et al., IEEE Trans. Power Electron., 18, 3, 907–914, 2003, DOI:
       10.1109/TPEL.2003.810840.
[34]   R. L. Coffie, “High Power High Frequency Transistors: A Material’s Perspective,” Springer
       International Publishing, Cham, 2020, 5–41.
[36] B. J. Baliga, IEEE Electron Device Lett., 10, 10, 455–457, 1989, DOI: 10.1109/55.43098.
[38]   P. Murugapandiyan et al., J. Sci. Adv. Mater. Devices, 2, 4, 515–522, 2017, DOI:
       10.1016/j.jsamd.2017.08.004.
[39] F. Yun et al., J. Appl. Phys., 92, 8, 4837–4839, 2002, DOI: 10.1063/1.1508420.
[40]   S. Kasap et al., Eds., Springer Handbook of Electronic and Photonic Materials. Cham:
       Springer International Publishing, 2017.
[41] I. Vurgaftman et al., J. Appl. Phys., 89, 11, 5815–5875, 2001, DOI: 10.1063/1.1368156.
[42]   Q. Guo et al., Jpn. J. Appl. Phys., 33, Part 1, No. 5A, 2453–2456, 1994, DOI:
       10.1143/JJAP.33.2453.
[44]   F. Bernardini et al., Phys. Rev. B, 56, 16, R10024–R10027, 1997, DOI:
       10.1103/PhysRevB.56.R10024.
[46]   M.   Posternak    et   al.,   Phys.   Rev.   Lett.,   64,   15,   1777–1780,   1990,   DOI:
       10.1103/PhysRevLett.64.1777.
[47] Semiconductor Device Physics and Design. Dordrecht: Springer Netherlands, 2007.
[48] O. Ambacher et al., J. Appl. Phys., 87, 1, 334–344, 2000, DOI: 10.1063/1.371866.
[49] Gallium Nitride Electronics, 96. Berlin, Heidelberg: Springer Berlin Heidelberg, 2008.
[50]   E. T. Yu et al., J. Vac. Sci. Technol. B Microelectron. Nanom. Struct., 17, 4, 1742, 1999,
       DOI: 10.1116/1.590818.
[51]   I. Vurgaftman et al., J. Appl. Phys., 94, 6, 3675–3696, 2003, DOI: 10.1063/1.1600519.
[52]   H. Morkoc et al., “Polarization in GaN Based Heterostructures and Heterojunction Field
       Effect Transistors (HFETs),” Springer US, Boston, MA, 2008, 373–466.
[53] G. D. O’Clock et al., Appl. Phys. Lett., 23, 2, 55–56, 1973, DOI: 10.1063/1.1654804.
[54] O. Ambacher et al., J. Appl. Phys., 85, 6, 3222–3233, 1999, DOI: 10.1063/1.369664.
[55] J. P. Ibbetson et al., Appl. Phys. Lett., 77, 2, 250–252, 2000, DOI: 10.1063/1.126940.
[56]   L. Gordon et al., J. Phys. D. Appl. Phys., 43, 50, 505501, 2010, DOI: 10.1088/0022-
       3727/43/50/505501.
[57] T. Zhu et al., Phys. Chem. Chem. Phys., 14, 27, 9558, 2012, DOI: 10.1039/c2cp40998d.
[60]   S.   Khandelwal     et   al.,    Solid.   State.     Electron.,    76,   60–66,    2012,   DOI:
       10.1016/j.sse.2012.05.054.
[61] S. Kola et al., IEEE Electron Device Lett., 9, 3, 136–138, 1988, DOI: 10.1109/55.2067.
[62]   D. Delagebeaudeuf et al., IEEE Trans. Electron Devices, 29, 6, 955–960, 1982, DOI:
       10.1109/T-ED.1982.20813.
[65]   “Design considerations of GaN devices for improving powerconverter efficiency and
       density.”                                 [Online].                                   Available:
       https://www.ti.com/lit/wp/slyy124/slyy124.pdf?ts=1598362572806&ref_url=https%253A
       %252F%252Fwww.google.com%252F.
[66]   T. Heckel et al., “Characterization and application of 600 V normally-off GaN transistors
       in hard switching DC/DC converters,” in 2014 IEEE 26th International Symposium on
       Power       Semiconductor       Devices   &        IC’s   (ISPSD),       2014,    63–66,   DOI:
       10.1109/ISPSD.2014.6855976.
[67]   K. J. Chen et al., IEEE Trans. Electron Devices, 64, 3, 779–795, 2017, DOI:
       10.1109/TED.2017.2657579.
[68]   X.-G. He et al., Chinese Phys. B, 24, 6, 067301, 2015, DOI: 10.1088/1674-
       1056/24/6/067301.
[70]   M. Haeberlen et al., J. Phys. Conf. Ser., 209, 012017, 2010, DOI: 10.1088/1742-
       6596/209/1/012017.
[71] D. Christy et al., AIP Adv., 4, 10, 107104, 2014, DOI: 10.1063/1.4897338.
[72]   L. Liu et al., Mater. Sci. Eng. R Reports, 37, 3, 61–127, 2002, DOI: 10.1016/S0927-
       796X(02)00008-6.
[73] R. Liu et al., Appl. Phys. Lett., 83, 5, 860–862, 2003, DOI: 10.1063/1.1597749.
[74] A. Tajalli et al., Materials (Basel)., 13, 19, 4271, 2020, DOI: 10.3390/ma13194271.
[75]   S. M. Cho et al., “AlGaN/GaN HFET grown on 6-inch diameter Si(111) substrates by
       MOCVD,” in Extended Abstracts of the 2011 International Conference on Solid State
       Devices and Materials, 2011, DOI: 10.7567/SSDM.2011.AL-7-3.
[76]   A.   Asgari   et   al.,   Mater.   Sci.   Eng.   C,   26,   5–7,   898–901,   2006,   DOI:
       10.1016/j.msec.2005.09.002.
[77]   X. Huang et al., “Evaluation and application of 600V GaN HEMT in cascode structure,” in
       2013 Twenty-Eighth Annual IEEE Applied Power Electronics Conference and Exposition
       (APEC), 2013, 1279–1286, DOI: 10.1109/APEC.2013.6520464.
[78]   X. Huang et al., IEEE Trans. Power Electron., 29, 5, 2208–2219, 2014, DOI:
       10.1109/TPEL.2013.2267804.
[79]   G. Sorrentino et al., “GaN HEMT devices: Experimental results on normally-on, normally-
       off and cascode configuration,” in IECON 2013 - 39th Annual Conference of the IEEE
       Industrial Electronics Society, 2013, 816–821, DOI: 10.1109/IECON.2013.6699239.
[80]   K. J. Chen et al., “Physics of fluorine plasma ion implantation for GaN normally-off HEMT
       technology,” in 2011 International Electron Devices Meeting, 2011, 19.4.1-19.4.4, DOI:
       10.1109/IEDM.2011.6131585.
[81]   Z. Zhang et al., IEEE Electron Device Lett., 36, 11, 1128–1131, 2015, DOI:
       10.1109/LED.2015.2483760.
[82]   C.-H. Wu et al., IEEE J. Electron Devices Soc., 6, 893–899, 2018, DOI:
       10.1109/JEDS.2018.2859769.
[83]   W. Saito et al., IEEE Trans. Electron Devices, 53, 2, 356–362, 2006, DOI:
       10.1109/TED.2005.862708.
[84]   S. D. Burnham et al., Phys. Status Solidi (c), 7, 7–8, 2010–2012, 2010, DOI:
       10.1002/pssc.200983644.
[85]   T. Oka et al., IEEE Electron Device Lett., 29, 7, 668–670, 2008, DOI:
       10.1109/LED.2008.2000607.
[86]   O. Hilt et al., “Normally-off high-voltage p-GaN gate GaN HFET with carbon-doped
       buffer,” in 2011 IEEE 23rd International Symposium on Power Semiconductor Devices and
       ICs, 2011, 239–242, DOI: 10.1109/ISPSD.2011.5890835.
[88]   P. Fiorenza et al., J. Vac. Sci. Technol. B, Nanotechnol. Microelectron. Mater. Process.
       Meas. Phenom., 35, 1, 01A101, 2017, DOI: 10.1116/1.4967306.
[89] G. Greco et al., Nanoscale Res. Lett., 6, 1, 132, 2011, DOI: 10.1186/1556-276X-6-132.
[90] S. C. Kang et al., Nanomaterials, 10, 11, 2116, 2020, DOI: 10.3390/nano10112116.
[92]   G. Greco et al., Mater. Sci. Semicond. Process., 78, 96–106, 2018, DOI:
       10.1016/j.mssp.2017.09.027.
[93]   A. Lorenz et al., Phys. Status Solidi (c), 6, S2, S996–S998, 2009, DOI:
       10.1002/pssc.200880838.
[94]   K. J. Chen et al., Phys. Status Solidi (a), 208, 2, 434–438, 2011, DOI:
       10.1002/pssa.201000631.
[96] T. Fujii et al., Jpn. J. Appl. Phys., 46, 1, 115–118, 2007, DOI: 10.1143/JJAP.46.115.
[97] T. Mori et al., Appl. Phys. Lett., 69, 23, 3537–3539, 1996, DOI: 10.1063/1.117237.
[98]   C. Auth et al., “A 22nm high performance and low-power CMOS technology featuring
       fully-depleted tri-gate transistors, self-aligned contacts and high density MIM capacitors,”
       in   2012   Symposium     on    VLSI   Technology     (VLSIT),    2012,   131–132,      DOI:
       10.1109/VLSIT.2012.6242496.
[99] T. Tamura et al., Appl. Phys. Express, 1, 023001, 2008, DOI: 10.1143/APEX.1.023001.
[100] K. Ohi et al., Jpn. J. Appl. Phys., 48, 8, 081002, 2009, DOI: 10.1143/JJAP.48.081002.
[101] J. Ma et       al., IEEE Electron Device Lett., 38, 3, 367–370,                2017, DOI:
       10.1109/LED.2017.2661755.
[102] J. Ma et al., IEEE Electron Device Lett., 38, 9, 1305–1308, 2017, DOI:
       10.1109/LED.2017.2731799.
[103] S. Takashima et al., IEEE Trans. Electron Devices, 60, 10, 3025–3031, 2013, DOI:
       10.1109/TED.2013.2278185.
[104] J. Ma et al., IEEE Trans. Electron Devices, 66, 9, 4068–4074, 2019, DOI:
       10.1109/TED.2019.2925859.
[105] M. Azize et al., Appl. Phys. Lett., 98, 4, 042103, 2011, DOI: 10.1063/1.3544048.
[106] O. Landré et al., Phys. Rev. B, 81, 15, 153306, 2010, DOI: 10.1103/PhysRevB.81.153306.
[107] K.-S. Im et al., IEEE Trans. Electron Devices, 60, 10, 3012–3018, 2013, DOI:
       10.1109/TED.2013.2274660.
[108] K. Ren et al., “Physical mechanism of fin-gate AlGaN/GaN MIS-HEMT: Vth model,” in
      2016 IEEE 4th Workshop on Wide Bandgap Power Devices and Applications (WiPDA),
      2016, 319–323, DOI: 10.1109/WiPDA.2016.7799960.
[109] C.-Y. Chien et al., Nanoscale Res. Lett., 12, 1, 420, 2017, DOI: 10.1186/s11671-017-2189-
      3.
[110] L. Nela et al., IEEE Electron Device Lett., 40, 3, 439–442, 2019, DOI:
      10.1109/LED.2019.2896359.
[111] B. Lu et al., IEEE Electron Device Lett., 33, 3, 360–362, 2012, DOI:
      10.1109/LED.2011.2179971.
[112] M. Zhu et al., IEEE Electron Device Lett., 40, 8, 1289–1292, 2019, DOI:
      10.1109/LED.2019.2922204.
[113] M.   Zhu    et    al.,     IEEE    Electron   Device    Lett.,    42,    1,     82–85,   2021,   DOI:
      10.1109/LED.2020.3037026.
[114] Y. Ma et al., Appl. Phys. Lett., 117, 14, 143506, 2020, DOI: 10.1063/5.0025351.
[115] M. Zhu et al., “Monolithic integration of GaN-based NMOS digital logic gate circuits with
      E-mode power GaN MOSHEMTs,” in 2018 IEEE 30th International Symposium on Power
      Semiconductor            Devices     and      ICs      (ISPSD),         2018,      236–239,      DOI:
      10.1109/ISPSD.2018.8393646.
[116] Y. Cai et al., IEEE Electron Device Lett., 28, 5, 328–331, 2007, DOI:
      10.1109/LED.2007.895391.
[117] G. Tang et al., IEEE Electron Device Lett., 38, 9, 1282–1285, 2017, DOI:
      10.1109/LED.2017.2725908.
[118] Z.   Xu    et    al.,     IEEE     Electron   Device    Lett.,    35,    1,     33–35,   2014,   DOI:
      10.1109/LED.2013.2291854.
[119] N. Chowdhury et al., IEEE Electron Device Lett., 40, 7, 1036–1039, 2019, DOI:
      10.1109/LED.2019.2916253.
[120] Z. Zheng et al., “Enhancement-Mode GaN p-Channel MOSFETs for Power Integration,” in
      2020 32nd International Symposium on Power Semiconductor Devices and ICs (ISPSD),
      2020, 525–528, DOI: 10.1109/ISPSD46842.2020.9170081.
[121] N. Chowdhury et al., IEEE Electron Device Lett., 41, 6, 820–823, 2020, DOI:
      10.1109/LED.2020.2987003.
[122] S. R. Bahl et al., “A New Approach to Validate GaN FET Reliability to Power-Line Surges
      Under Use-Conditions,” in 2019 IEEE International Reliability Physics Symposium (IRPS),
      2019, 1–4, DOI: 10.1109/IRPS.2019.8720479.
[123] Z. H. Liu et al., Appl. Phys. Lett., 98, 11, 113506, 2011, DOI: 10.1063/1.3567927.
[124] S. Kato et al., J. Cryst. Growth, 298, 831–834, 2007, DOI: 10.1016/j.jcrysgro.2006.10.192.
[125] Y. C. Choi et al., Semicond. Sci. Technol., 22, 5, 517–521, 2007, DOI: 10.1088/0268-
      1242/22/5/010.
[126] Y. C. Choi et al., “Effects of an Fe-doped GaN Buffer in AlGaN/GaN Power HEMTs on Si
      Substrate,” in 2006 European Solid-State Device Research Conference, 2006, 282–285,
      DOI: 10.1109/ESSDER.2006.307693.
[127] F.   Medjdoub     et   al.,   Appl.   Phys.   Express,   4,   12,   124101,   2011,    DOI:
      10.1143/APEX.4.124101.
[128] I. Abid et al., Micromachines, 10, 10, 690, 2019, DOI: 10.3390/mi10100690.
[129] J.-H. Lee et al., IEEE J. Electron Devices Soc., 6, 1179–1186, 2018, DOI:
      10.1109/JEDS.2018.2872975.
[130] Xinhua Wang et al., IEEE Trans. Electron Devices, 61, 5, 1341–1346, 2014, DOI:
      10.1109/TED.2014.2312232.
[131] X.-H. Ma et al., Chinese Phys. B, 24, 2, 027101, 2015, DOI: 10.1088/1674-
      1056/24/2/027101.
[132] S. Besendörfer et al., J. Appl. Phys., 127, 1, 015701, 2020, DOI: 10.1063/1.5129248.
[133] G. Meneghesso et al., Jpn. J. Appl. Phys., 53, 10, 100211, 2014, DOI:
      10.7567/JJAP.53.100211.
[134] R.   Coffie,   IEEE    Trans.   Electron   Devices,   61,       3,        878–883,   2014,   DOI:
      10.1109/TED.2014.2300115.
[135] S. Karmalkar et al., IEEE Trans. Electron Devices, 52, 12, 2534–2540, 2005, DOI:
      10.1109/TED.2005.859568.
[136] R. Coffie, IEEE Trans. Electron            Devices, 61, 8, 2867–2872, 2014, DOI:
      10.1109/TED.2014.2329475.
[137] S. Karmalkar et al., IEEE Trans. Electron Devices, 48, 8, 1515–1521, 2001, DOI:
      10.1109/16.936500.
[138] Y. Dora et al., IEEE Electron Device Lett., 27, 9, 713–715, 2006, DOI:
      10.1109/LED.2006.881020.
[139] I. Rossetto et al., IEEE Trans. Electron Devices, 64, 1, 73–77, 2017, DOI:
      10.1109/TED.2016.2623774.
[140] V. Kumar et al., IEEE Trans. Electron Devices, 53, 6, 1477–1480, 2006, DOI:
      10.1109/TED.2006.874090.
[141] Rongming Chu et al., IEEE Electron Device Lett., 32, 5, 632–634, 2011, DOI:
      10.1109/LED.2011.2118190.
[142] J. Ma et al., Appl. Phys. Lett., 112, 5, 052101, 2018, DOI: 10.1063/1.5012866.
[143] E. Matioli et al., IEEE Trans. Electron Devices, 60, 10, 3365–3370, 2013, DOI:
      10.1109/TED.2013.2279120.
[144] J.   Ma   et   al.,   IEEE   Electron   Device   Lett.,   38,        1,     83–86,   2017,   DOI:
      10.1109/LED.2016.2632044.
[145] L. Nela et al., IEEE Electron Device Lett., 41, 1, 99–102, 2020, DOI:
      10.1109/LED.2019.2957700.
[146] L. Nela et al., IEEE Trans. Power Electron., 36, 2, 1269–1273, 2021, DOI:
      10.1109/TPEL.2020.3008226.
[147] K.-S. Im et al., “Novel AlGaN/GaN omega-FinFETs with excellent device performances,”
      in 2016 46th European Solid-State Device Research Conference (ESSDERC), 2016, 323–
      326, DOI: 10.1109/ESSDERC.2016.7599651.
[148] K.-S. Im et al., Solid. State. Electron., 129, 196–199, 2017, DOI: 10.1016/j.sse.2016.11.005.
[149] Y. Xu et al., IEEE Trans. Electron Devices, 65, 3, 915–920, 2018, DOI:
      10.1109/TED.2017.2788920.
[150] Y. C. Choi et al., IEEE Trans. Electron Devices, 53, 12, 2926–2931, 2006, DOI:
      10.1109/TED.2006.885679.
[151] M. Meneghini et al., IEEE Trans. Electron Devices, 61, 12, 4070–4077, 2014, DOI:
      10.1109/TED.2014.2364855.
[152] G. Verzellesi et al., IEEE Electron Device Lett., 35, 4, 443–445, 2014, DOI:
      10.1109/LED.2014.2304680.
[153] R. Kabouche et al., Phys. Status Solidi (a), 217, 7, 1900687, 2020, DOI:
      10.1002/pssa.201900687.
[154] P. Srivastava et al., IEEE Electron Device Lett., 31, 8, 851–853, 2010, DOI:
      10.1109/LED.2010.2050673.
[155] N. Herbecq et al., Electron. Lett., 51, 19, 1532–1534, 2015, DOI: 10.1049/el.2015.1684.
[156] G. Pavlidis et al., IEEE Electron Device Lett., 40, 7, 1060–1063, 2019, DOI:
      10.1109/LED.2019.2915984.
[157] G. Pavlidis et al., “The thermal effects of substrate removal on GaN HEMTs using Raman
      Thermometry,” in Proceedings of the 15th InterSociety Conference on Thermal and
      Thermomechanical Phenomena in Electronic Systems, ITherm 2016, 2016, 1255–1260,
      DOI: 10.1109/ITHERM.2016.7517691.
[158] A. G. Baca et al., J. Vac. Sci. Technol. A, 38, 2, 020803, 2020, DOI: 10.1116/1.5129803.
[159] R. J. Kaplar et al., “III-Nitride ultra-wide-bandgap electronic devices,” Academic Press Inc.,
      2019, 397–416.
[160] T. Razzak et al., Int. J. High Speed Electron. Syst., 28, 1–2, 2019, DOI:
      10.1142/S0129156419400093.
[162] E. A. Douglas et al., Phys. Status Solidi (a), 214, 8, 1600842, 2017, DOI:
      10.1002/pssa.201600842.
[163] N. Yafune et al., Jpn. J. Appl. Phys., 50, 10 PART 1, 100202, 2011, DOI:
      10.1143/JJAP.50.100202.
[165] S. Bajaj et al., Appl. Phys. Lett., 109, 13, 133508, 2016, DOI: 10.1063/1.4963860.
[167] B. J. Baliga, Semicond. Sci. Technol., 28, 7, 074011, 2013, DOI: 10.1088/0268-
      1242/28/7/074011.
[169] S. Heikman et al., J. Appl. Phys., 94, 8, 5321, 2003, DOI: 10.1063/1.1610244.
[170] T. Palacios et al., IEEE Trans. Electron Devices, 53, 3, 562–565, 2006, DOI:
      10.1109/TED.2005.863767.
[171] J. Lei et        al., IEEE Electron Device Lett., 39, 2, 260–263, 2018, DOI:
      10.1109/LED.2017.2783908.
[172] R. S. Howell et al., “The Super-Lattice Castellated Field Effect Transistor (SLCFET): A
      novel high performance Transistor topology ideal for RF switching,” in 2014 IEEE
      International        Electron         Devices      Meeting,      2014,     11.5.1-11.5.4,   DOI:
      10.1109/IEDM.2014.7047033.
[173] K. Shinohara et al., “GaN-Based Multi-Channel Transistors with Lateral Gate for Linear
      and Efficient Millimeter-Wave Power Amplifiers,” in 2019 IEEE MTT-S International
      Microwave Symposium (IMS), 2019, 1133–1135, DOI: 10.1109/MWSYM.2019.8700845.
[174] J. Chang et al., IEEE Electron Device Lett., 40, 7, 1048–1051, 2019, DOI:
      10.1109/LED.2019.2917285.
[175] C. Erine et al., IEEE Electron Device Lett., 41, 3, 321–324, 2020, DOI:
      10.1109/LED.2020.2967458.
[176] J. Ma et al., Appl. Phys. Lett., 113, 24, 242102, 2018, DOI: 10.1063/1.5064407.
[177] J. Ma et         al., IEEE Electron Device Lett., 40, 2, 275–278, 2019, DOI:
      10.1109/LED.2018.2887199.
[178] J. Ma et al., “1200 V Multi-Channel Power Devices with 2.8 Ω•mm ON-Resistance,” in
      2019 IEEE International Electron Devices Meeting (IEDM), 2019, 4.1.1-4.1.4, DOI:
      10.1109/IEDM19573.2019.8993536.
[179] M. Xiao et al., IEEE Electron Device Lett., 41, 8, 1177–1180, 2020, DOI:
      10.1109/LED.2020.3005934.
[180] L.   Nela   et    al.,   IEEE   Electron   Device   Lett.,   42,   1,   86–89,   2021,   DOI:
      10.1109/LED.2020.3038808.
[181] H. Kawai et al., Phys. Status Solidi (a), 214, 8, 1600834, 2017, DOI:
      10.1002/pssa.201600834.
[182] T. Fujihira, Jpn. J. Appl. Phys., 36, Part 1, No. 10, 6254–6262, 1997, DOI:
      10.1143/JJAP.36.6254.
[183] F. Udrea et al., IEEE Trans. Electron Devices, 64, 3, 713–727, 2017, DOI:
      10.1109/TED.2017.2658344.
[184] X. Zhou et al., Appl. Phys. Lett., 115, 11, 112104, 2019, DOI: 10.1063/1.5109389.
[185] H. Huang et al., Appl. Phys. Lett., 116, 10, 102103, 2020, DOI: 10.1063/1.5142855.
[186] Z. Li et al., IEEE Trans. Electron Devices, 60, 10, 3230–3237, 2013, DOI:
      10.1109/TED.2013.2266544.
[187] H. Ishida et al., IEEE Electron Device Lett., 29, 10, 1087–1089, 2008, DOI:
      10.1109/LED.2008.2002753.
[188] H. Ishida et al., “GaN-based natural super junction diodes with multi-channel structures,”
      in   2008    IEEE   International      Electron    Devices    Meeting,     2008,    1–4,   DOI:
      10.1109/IEDM.2008.4796636.
[189] A. Nakajima et al., IEEE Electron Device Lett., 32, 4, 542–544, 2011, DOI:
      10.1109/LED.2011.2105242.
[190] B. Song et al., “Design and optimization of GaN lateral polarization-doped super-junction
      (LPSJ): An analytical study,” in 2015 IEEE 27th International Symposium on Power
      Semiconductor       Devices     &        IC’s      (ISPSD),     2015,       273–276,       DOI:
      10.1109/ISPSD.2015.7123442.
[191] S.-W. Han et al., IEEE Trans. Electron Devices, 67, 1, 69–74, 2020, DOI:
      10.1109/TED.2019.2953843.
[192] S.-W. Han et al., IEEE Electron Device Lett., 41, 12, 1758–1761, 2020, DOI:
      10.1109/LED.2020.3029619.
[195] “Product Selector Guide for eGaN® FETs and ICs.” [Online]. Available: https://epc-
      co.com/epc/Products/eGaNFETsandICs.aspx.
[196] J. A. del Alamo et al., Microelectron. Reliab., 49, 9–11, 1200–1206, 2009, DOI:
      10.1016/j.microrel.2009.07.003.
[197] T. Ohki et al., “Reliability of GaN HEMTs: current status and future technology,” in 2009
      IEEE     International   Reliability     Physics     Symposium,     2009,       61–70,     DOI:
      10.1109/IRPS.2009.5173225.
[198] W. Saito, “Reliability of GaN-HEMTs for high-voltage switching applications,” in 2011
      International   Reliability     Physics   Symposium,       2011,   4E.1.1-4E.1.5,    DOI:
      10.1109/IRPS.2011.5784510.
[199] G. Meneghesso et al., IEEE Trans. Device Mater. Reliab., 8, 2, 332–343, 2008, DOI:
      10.1109/TDMR.2008.923743.
[200] W. Zhang et al., “Evaluation of 600 V cascode GaN HEMT in device characterization and
      all-GaN-based LLC resonant converter,” in 2013 IEEE Energy Conversion Congress and
      Exposition, 2013, 3571–3578, DOI: 10.1109/ECCE.2013.6647171.
[201] X. Huang et al., IEEE Trans. Power Electron., 29, 5, 2453–2461, 2014, DOI:
      10.1109/TPEL.2013.2276127.
[202] Z. Liu et al., IEEE Trans. Power Electron., 29, 4, 1977–1985, 2014, DOI:
      10.1109/TPEL.2013.2264941.
[203] Yong Cai et al., IEEE Electron Device Lett., 26, 7, 435–437, 2005, DOI:
      10.1109/LED.2005.851122.
[205] Y. Zhang et al., Appl. Phys. Lett., 111, 16, 163506, 2017, DOI: 10.1063/1.4989599.
[206] K. Mochizuki et al.,          Jpn. J. Appl. Phys., 52, 8S, 08JN22, 2013, DOI:
      10.7567/JJAP.52.08JN22.
[207] B. J. Baliga, Fundamentals of Power Semiconductor Devices. Boston, MA: Springer US,
      2008.
[208] I. C. Kizilyalli et al., IEEE Trans. Electron Devices, 60, 10, 3067–3070, 2013, DOI:
      10.1109/TED.2013.2266664.
[209] C. De Santi et al., IEEE Electron Device Lett., 41, 9, 1300–1303, 2020, DOI:
      10.1109/LED.2020.3009649.
[210] T. Maeda et al., IEEE Electron Device Lett., 40, 6, 941–944, 2019, DOI:
      10.1109/LED.2019.2912395.
[211] R. Kucharski et al., J. Appl. Phys., 128, 5, 050902, 2020, DOI: 10.1063/5.0009900.
[212] S. T. Kim et al., J. Cryst. Growth, 194, 1, 37–42, 1998, DOI: 10.1016/S0022-
      0248(98)00551-X.
[214] H. Yamane et al., J. Cryst. Growth, 186, 1–2, 8–12, 1998, DOI: 10.1016/S0022-
      0248(97)00480-6.
[215] R.   Dwiliński    et   al.,   J.   Cryst.   Growth, 311,   10,   3015–3018,    2009,   DOI:
      10.1016/j.jcrysgro.2009.01.052.
[216] R.   Dwiliński    et   al.,   J.   Cryst.   Growth, 310,   17,   3911–3916,    2008,   DOI:
      10.1016/j.jcrysgro.2008.06.036.
[217] H. Ohta et al., Jpn. J. Appl. Phys., 57, 4S, 04FG09, 2018, DOI: 10.7567/JJAP.57.04FG09.
[218] H. Ohta et al., IEEE Electron Device Lett., 36, 11, 1180–1182, 2015, DOI:
      10.1109/LED.2015.2478907.
[219] Y. Zhang et al., J. Phys. D. Appl. Phys., 51, 27, 273001, 2018, DOI: 10.1088/1361-
      6463/aac8aa.
[222] “12 Inch Silicon Wafers for Research & Development.” [Online]. Available:
      https://www.universitywafer.com/12-inch-silicon-wafers.html.
[224] S. Raghavan et al., J. Appl. Phys., 98, 2, 023514, 2005, DOI: 10.1063/1.1978991.
[225] M. Miyoshi et al., Semicond. Sci. Technol., 31, 10, 105016, 2016, DOI: 10.1088/0268-
      1242/31/10/105016.
[226] S. Chowdhury, “GaN-on-GaN power device design and fabrication,” Elsevier, 2019, 209–
      248.
[227] I. C. Kizilyalli et al., Microelectron. Reliab., 55, 9–10, 1654–1661, 2015, DOI:
      10.1016/j.microrel.2015.07.012.
[228] I. C. Kizilyalli et al., IEEE Trans. Electron Devices, 62, 2, 414–422, 2015, DOI:
      10.1109/TED.2014.2360861.
[229] Y. Hatakeyama et al., IEEE Electron Device Lett., 32, 12, 1674–1676, 2011, DOI:
      10.1109/LED.2011.2167125.
[230] J.-I. Chyi et al., Solid. State. Electron., 44, 4, 613–617, 2000, DOI: 10.1016/S0038-
      1101(99)00183-5.
[231] A. . Zhang et al., Solid. State. Electron., 44, 7, 1157–1161, 2000, DOI: 10.1016/S0038-
      1101(00)00059-9.
[232] X. A. Cao et al., Appl. Phys. Lett., 87, 5, 053503, 2005, DOI: 10.1063/1.2001738.
[233] K. Nomoto et al., Phys. Status Solidi (a), 208, 7, 1535–1537, 2011, DOI:
      10.1002/pssa.201000976.
[234] I. C. Kizilyalli et al., IEEE Electron Device Lett., 35, 2, 247–249, 2014, DOI:
      10.1109/LED.2013.2294175.
[235] Y. Hatakeyama et al., Jpn. J. Appl. Phys., 52, 2R, 028007, 2013, DOI:
      10.7567/JJAP.52.028007.
[236] K. Fu et al., Appl. Phys. Lett., 113, 23, 233502, 2018, DOI: 10.1063/1.5052479.
[237] H. Xing et al., J. Appl. Phys., 97, 11, 113703, 2005, DOI: 10.1063/1.1914952.
[238] C. Matthews et al., “Switching characterization of vertical GaN PiN diodes,” in 2016 IEEE
      4th Workshop on Wide Bandgap Power Devices and Applications (WiPDA), 2016, 135–
      138, DOI: 10.1109/WiPDA.2016.7799924.
[239] “Vertical Devices In Bulk GaN Drive Diode Performance To Near - Theoretical Limits.”
      [Online]. Available: http://www.how2power.com/article/2013/march/vertical-devices-in-
      bulk-gan-drive-diode-performance-to-near---theoretical-l-imits-.php.
[240] I. C. Kizilyalli et al., IEEE Electron Device Lett., 35, 6, 654–656, 2014, DOI:
      10.1109/LED.2014.2319214.
[241] Z. Hu et al., Appl. Phys. Lett., 107, 24, 243501, 2015, DOI: 10.1063/1.4937436.
[242] K. Nomoto et al., “GaN-on-GaN p-n power diodes with 3.48 kV and 0.95 mΩ-cm2: A
      record high figure-of-merit of 12.8 GW/cm2,” in 2015 IEEE International Electron Devices
      Meeting (IEDM), 2015, 9.7.1-9.7.4, DOI: 10.1109/IEDM.2015.7409665.
[243] Z. Hu et al., IEEE Electron Device Lett., 38, 8, 1071–1074, 2017, DOI:
      10.1109/LED.2017.2720747.
[244] M. Qi et al., Appl. Phys. Lett., 107, 23, 232101, 2015, DOI: 10.1063/1.4936891.
[245] R. Yeluri et al., Appl. Phys. Lett., 106, 18, 183502, 2015, DOI: 10.1063/1.4919866.
[246] H. Fu et al., IEEE Electron Device Lett., 39, 7, 1018–1021, 2018, DOI:
      10.1109/LED.2018.2837625.
[247] H. Fu et al., IEEE Electron Device Lett., 38, 6, 763–766, 2017, DOI:
      10.1109/LED.2017.2690974.
[248] K. Fu et al., IEEE Electron Device Lett., 40, 11, 1728–1731, 2019, DOI:
      10.1109/LED.2019.2941830.
[249] J. Wang et al., IEEE Electron Device Lett., 39, 11, 1716–1719, 2018, DOI:
      10.1109/LED.2018.2868560.
[250] J. Wang et al., Appl. Phys. Lett., 113, 2, 023502, 2018, DOI: 10.1063/1.5035267.
[251] J. R. Dickerson et al., IEEE Trans. Electron Devices, 63, 1, 419–425, 2016, DOI:
      10.1109/TED.2015.2502186.
[252] J. Lutz et al., “pin Diodes,” Springer International Publishing, Cham, 2018, 201–270.
[253] G. T. Dang et al., IEEE Trans. Electron Devices, 47, 4, 692–696, 2000, DOI:
      10.1109/16.830981.
[254] M. Misra et al., Appl. Phys. Lett., 76, 8, 1045–1047, 2000, DOI: 10.1063/1.125933.
[255] Y. Saitoh et al., Appl. Phys. Express, 3, 8, 081001, 2010, DOI: 10.1143/APEX.3.081001.
[256] M. Ueno et al., “Fast recovery performance of vertical GaN Schottky barrier diodes on low-
      dislocation-density GaN substrates,” in 2014 IEEE 26th International Symposium on Power
      Semiconductor      Devices      &     IC’s     (ISPSD),     2014,      309–312,     DOI:
      10.1109/ISPSD.2014.6856038.
[257] N. Tanaka et al., Appl. Phys. Express, 8, 7, 071001, 2015, DOI: 10.7567/APEX.8.071001.
[258] Y. Cao et al., Appl. Phys. Lett., 108, 11, 112101, 2016, DOI: 10.1063/1.4943946.
[259] Y. Cao et al., Appl. Phys. Lett., 108, 6, 062103, 2016, DOI: 10.1063/1.4941814.
[260] L. Sang et al., Appl. Phys. Lett., 111, 12, 122102, 2017, DOI: 10.1063/1.4994627.
[261] K. Nagamatsu et al., Jpn. J. Appl. Phys., 57, 10, 105501, 2018, DOI:
      10.7567/JJAP.57.105501.
[263] S. Abhinay et al., Jpn. J. Appl. Phys., 59, 1, 010906, 2020, DOI: 10.7567/1347-
      4065/ab65cd.
[264] H. Fu et al., Appl. Phys. Lett., 111, 15, 152102, 2017, DOI: 10.1063/1.4993201.
[265] S. Han et al., IEEE Electron Device Lett., 39, 4, 572–575, 2018, DOI:
      10.1109/LED.2018.2808684.
[266] S. Han et al., IEEE Trans. Power Electron., 34, 6, 5012–5018, 2019, DOI:
      10.1109/TPEL.2018.2876444.
[267] S. Han et al., IEEE Electron Device Lett., 40, 7, 1040–1043, 2019, DOI:
      10.1109/LED.2019.2915578.
[268] W.-F. Wang et al., Chinese Phys. B, 29, 4, 047305, 2020, DOI: 10.1088/1674-1056/ab7909.
[269] W. Li et al., IEEE Trans. Electron Devices, 64, 4, 1635–1641, 2017, DOI:
      10.1109/TED.2017.2662702.
[270] Y. Zhang et al., IEEE Electron Device Lett., 38, 8, 1097–1100, 2017, DOI:
      10.1109/LED.2017.2720689.
[271] T.   Hayashida         et   al.,    Appl.     Phys.   Express,    10,    6,     061003,   2017,    DOI:
      10.7567/APEX.10.061003.
[272] A. D. Koehler et al., ECS J. Solid State Sci. Technol., 6, 1, Q10–Q12, 2017, DOI:
      10.1149/2.0041701jss.
[273] Y. Zhang et al., “Novel GaN trench MIS barrier Schottky rectifiers with implanted field
      rings,” in 2016 IEEE International Electron Devices Meeting (IEDM), 2016, 10.2.1-10.2.4,
      DOI: 10.1109/IEDM.2016.7838386.
[274] S. Chowdhury et al., IEEE Electron Device Lett., 29, 6, 543–545, 2008, DOI:
      10.1109/LED.2008.922982.
[275] D. Shibata et al., “1.7 kV/1.0 mΩcm 2 normally-off vertical GaN transistor on GaN
      substrate with regrown p-GaN/AlGaN/GaN semipolar gate structure,” in 2016 IEEE
      International      Electron        Devices     Meeting    (IEDM),       2016,     10.1.1-10.1.4,   DOI:
      10.1109/IEDM.2016.7838385.
[276] I. Ben-Yaacov et al., J. Appl. Phys., 95, 4, 2073–2078, 2004, DOI: 10.1063/1.1641520.
[277] T. J. Rodgers et al., IEEE J. Solid-State Circuits, 10, 5, 322–331, 1975, DOI:
      10.1109/JSSC.1975.1050618.
[278] S. Chowdhury et al., IEEE Electron Device Lett., 33, 1, 41–43, 2012, DOI:
      10.1109/LED.2011.2173456.
[279] H. Nie et al., IEEE Electron Device Lett., 35, 9, 939–941, 2014, DOI:
      10.1109/LED.2014.2339197.
[280] D. Ji et al., IEEE Trans. Electron Devices, 64, 3, 805–808, 2017, DOI:
      10.1109/TED.2016.2632150.
[281] D.   Ji   et    al.,    IEEE       Electron    Device    Lett.,   39,   6,    863–865,    2018,    DOI:
      10.1109/LED.2018.2828844.
[282] T. Oka et al., Appl. Phys. Express, 8, 5, 054101, 2015, DOI: 10.7567/APEX.8.054101.
[283] T. Oka et al., “Over 10 a operation with switching characteristics of 1.2 kV-class vertical
      GaN trench MOSFETs on a bulk GaN substrate,” in 2016 28th International Symposium on
      Power     Semiconductor        Devices     and   ICs    (ISPSD),    2016,      459–462,     DOI:
      10.1109/ISPSD.2016.7520877.
[284] H. Otake et al., Jpn. J. Appl. Phys., 46, No. 25, L599–L601, 2007, DOI:
      10.1143/JJAP.46.L599.
[285] H. Otake et al., Appl. Phys. Express, 1, 1, 011105, 2008, DOI: 10.1143/APEX.1.011105.
[286] M. Kodama et al., Appl. Phys. Express, 1, 021104, 2008, DOI: 10.1143/APEX.1.021104.
[287] T. Oka et al., Appl. Phys. Express, 7, 2, 021002, 2014, DOI: 10.7567/APEX.7.021002.
[288] T. Oka et al., “100 A Vertical GaN Trench MOSFETs with a Current Distribution Layer,”
      in 2019 31st International Symposium on Power Semiconductor Devices and ICs (ISPSD),
      2019, 303–306, DOI: 10.1109/ISPSD.2019.8757621.
[290] C. Gupta et al., IEEE Electron Device Lett., 38, 3, 353–355, 2017, DOI:
      10.1109/LED.2017.2649599.
[291] D. Ji et al., “Demonstrating >1.4 kV OG-FET performance with a novel double field-
      plated geometry and the successful scaling of large-area devices,” in 2017 IEEE
      International     Electron     Devices     Meeting     (IEDM),     2017,     9.4.1-9.4.4,   DOI:
      10.1109/IEDM.2017.8268359.
[292] C. Gupta et al., IEEE Electron Device Lett., 37, 12, 1601–1604, 2016, DOI:
      10.1109/LED.2016.2616508.
[293] W. Li et al., IEEE Trans. Electron Devices, 65, 6, 2558–2564, 2018, DOI:
      10.1109/TED.2018.2829125.
[294] W. Li et al., “600 V GaN vertical V-trench MOSFET with MBE regrown channel,” in 2017
      75th    Annual      Device      Research     Conference      (DRC),        2017,   1–2,     DOI:
      10.1109/DRC.2017.7999414.
[295] Y. Zhang et al., “1200 V GaN vertical fin power field-effect transistors,” in 2017 IEEE
      International   Electron     Devices   Meeting    (IEDM),    2017,    9.2.1-9.2.4,   DOI:
      10.1109/IEDM.2017.8268357.
[296] M. Sun et al., IEEE Electron Device Lett., 38, 4, 509–512, 2017, DOI:
      10.1109/LED.2017.2670925.
[297] M. Ruzzarin et al., IEEE Trans. Electron Devices, 64, 8, 3126–3131, 2017, DOI:
      10.1109/TED.2017.2716982.
[298] Y. Zhang et al., IEEE Electron Device Lett., 1–1, 2018, DOI: 10.1109/LED.2018.2880306.
[300] K. Strempel et al., Semicond. Sci. Technol., 36, 1, 014002, 2021, DOI: 10.1088/1361-
      6641/abc5ff.
[301] M. Ruzzarin et al., Appl. Phys. Lett., 117, 20, 203501, 2020, DOI: 10.1063/5.0027922.
[302] IEEE Electron Device Lett., 35, 6, 618–620, 2014, DOI: 10.1109/LED.2014.2314637.
[303] Yuhao Zhang et al., IEEE Trans. Electron Devices, 62, 7, 2155–2161, 2015, DOI:
      10.1109/TED.2015.2426711.
[305] H. Amano et al., J. Phys. D. Appl. Phys., 51, 16, 163001, 2018, DOI: 10.1088/1361-
      6463/aaaf9d.
[306] A. Magnani et al., “Thermal resistance characterization of GaN power HEMTs on Si, SOI,
      and poly-AlN substrates,” in 2020 21st International Conference on Thermal, Mechanical
      and Multi-Physics Simulation and Experiments in Microelectronics and Microsystems
      (EuroSimE), 2020, 1–6, DOI: 10.1109/EuroSimE48426.2020.9152656.
[308] X. Zou et al., IEEE Electron Device Lett., 37, 5, 636–639, 2016, DOI:
      10.1109/LED.2016.2548488.
[309] Y. Zhang et al., IEEE Electron Device Lett., 38, 2, 248–251, 2017, DOI:
      10.1109/LED.2016.2646669.
[310] S. Mase et al., Appl. Phys. Express, 9, 11, 111005, 2016, DOI: 10.7567/APEX.9.111005.
[311] Y. Zhang et al., IEEE Electron Device Lett., 39, 5, 715–718, 2018, DOI:
      10.1109/LED.2018.2819642.
[312] R. Abdul Khadar et al., IEEE Electron Device Lett., 39, 3, 401–404, 2018, DOI:
      10.1109/LED.2018.2793669.
[313] C.   Liu   et   al.,   IEEE   Electron   Device   Lett.,   39,   1,   71–74,   2018,   DOI:
      10.1109/LED.2017.2779445.
[314] R. A. Khadar et al., IEEE Electron Device Lett., 40, 3, 443–446, 2019, DOI:
      10.1109/LED.2019.2894177.
[315] W. Chen et al., “Monolithic integration of lateral field-effect rectifier with normally-off
      HEMT for GaN-on-Si switch-mode power supply converters,” in 2008 IEEE International
      Electron Devices Meeting, 2008, 1–4, DOI: 10.1109/IEDM.2008.4796635.
[316] P. Chyurlia et al., Phys. Status Solidi (c), 8, 7–8, 2210–2212, 2011, DOI:
      10.1002/pssc.201000914.
[317] W. Shaobing et al., IEEE Trans. Electron Devices, 63, 10, 3882–3886, 2016, DOI:
      10.1109/TED.2016.2597244.
[318] J. Ma et al., IEEE Electron Device Lett., 38, 12, 1704–1707, 2017, DOI:
      10.1109/LED.2017.2761911.
[319] C. Liu et al., IEEE Electron Device Lett., 39, 7, 1034–1037, 2018, DOI:
      10.1109/LED.2018.2841959.
[320] K. Mukherjee et al., Materials (Basel)., 13, 21, 4740, 2020, DOI: 10.3390/ma13214740.
[321] K. Mukherjee et al., Appl. Phys. Express, 13, 2, 024004, 2020, DOI: 10.35848/1882-
      0786/ab6ddd.
[322] “JC-70 Wide Bandgap Power Electronic Conversion Semiconductors.” [Online]. Available:
      https://www.jedec.org/committees/jc-70.
[325] X. She et al., IEEE Trans. Ind. Electron., 64, 10, 8193–8205, 2017, DOI:
      10.1109/TIE.2017.2652401.
[326] M. J. Tadjer et al., ECS J. Solid State Sci. Technol., 5, 2, P124–P127, 2016, DOI:
      10.1149/2.0371602jss.
[327] M. Bockowski, Appl. Phys. Lett., 115, 14, 142104, 2019, DOI: 10.1063/1.5116866.
[328] G. Koblmüller et al., J. Appl. Phys., 107, 4, 043527, 2010, DOI: 10.1063/1.3285309.
[329] W. Götz et al., Appl. Phys. Lett., 67, 18, 2666–2668, 1995, DOI: 10.1063/1.114330.
[330] W. Götz et al., Appl. Phys. Lett., 69, 24, 3725–3727, 1996, DOI: 10.1063/1.117202.
[331] S. Nakamura et al., Jpn. J. Appl. Phys., 31, Part 1, No. 5A, 1258–1266, 1992, DOI:
      10.1143/JJAP.31.1258.
[332] S. Nakamura et al., Jpn. J. Appl. Phys., 31, Part 2, No. 2B, L139–L142, 1992, DOI:
      10.1143/JJAP.31.L139.
[333] J.-D.        Hwang   et    al.,   Appl.     Surf.   Sci.,   253,    10,        4694–4697,   2007,   DOI:
      10.1016/j.apsusc.2006.10.026.
[334] C. Gupta et al., Appl. Phys. Express, 9, 12, 121001, 2016, DOI: 10.7567/APEX.9.121001.
[338] A. R. Peaker et al., J. Appl. Phys., 123, 16, 161559, 2018, DOI: 10.1063/1.5011327.
[339] R. Vetury et al., IEEE Trans. Electron Devices, 48, 3, 560–566, 2001, DOI:
      10.1109/16.906451.
[340] I. Daumiller et al., IEEE Electron Device Lett., 22, 2, 62–64, 2001, DOI:
      10.1109/55.902832.
[341] S. C. Binari et al., IEEE Trans. Electron Devices, 48, 3, 465–471, 2001, DOI:
      10.1109/16.906437.
[342] G. Meneghesso et al., IEEE Trans. Electron Devices, 53, 12, 2932–2941, 2006, DOI:
      10.1109/TED.2006.885681.
[343] M. Faqir et al., IEEE Trans. Device Mater. Reliab., 8, 2, 240–247, 2008, DOI:
      10.1109/TDMR.2008.922017.
[344] W. Saito et al., IEEE Trans. Electron Devices, 54, 8, 1825–1830, 2007, DOI:
      10.1109/TED.2007.901150.
[345] D. Jin et al., IEEE Trans. Electron Devices, 60, 10, 3190–3196, 2013, DOI:
      10.1109/TED.2013.2274477.
[346] D. Bisi et al., IEEE Electron Device Lett., 35, 10, 1004–1006, 2014, DOI:
      10.1109/LED.2014.2344439.
[347] M. Meneghini et al., IEEE Trans. Electron Devices, 62, 3, 782–787, 2015, DOI:
      10.1109/TED.2014.2386391.
[348] S. Huang et al., Jpn. J. Appl. Phys., 50, 110202, 2011, DOI: 10.1143/JJAP.50.110202.
[349] P. Lagger et al., “Towards understanding the origin of threshold voltage instability of
      AlGaN/GaN MIS-HEMTs,” in 2012 International Electron Devices Meeting, 2012, 13.1.1-
      13.1.4, DOI: 10.1109/IEDM.2012.6479033.
[350] M. Ťapajna et al., Appl. Phys. Lett., 102, 24, 243509, 2013, DOI: 10.1063/1.4811754.
[351] P. Lagger et al., IEEE Electron Device Lett., 34, 9, 1112–1114, 2013, DOI:
      10.1109/LED.2013.2272095.
[352] M. Van Hove et al., IEEE Trans. Electron Devices, 60, 10, 3071–3078, 2013, DOI:
      10.1109/TED.2013.2274730.
[353] M. Meneghini et al., IEEE Electron Device Lett., 37, 4, 474–477, 2016, DOI:
      10.1109/LED.2016.2530693.
[354] A. Guo et al., IEEE Trans. Electron Devices, 64, 5, 2142–2147, 2017, DOI:
      10.1109/TED.2017.2686840.
[355] G. Meneghesso et al., IEEE Electron Device Lett., 30, 2, 100–102, 2009, DOI:
      10.1109/LED.2008.2010067.
[356] M. Wang et al., IEEE Electron Device Lett., 32, 4, 482–484, 2011, DOI:
      10.1109/LED.2011.2105460.
[357] M.    Asghar    et    al.,   Mater.    Sci.    Eng.   B,   113,   3,   248–252,   2004,   DOI:
      10.1016/j.mseb.2004.09.001.
[358] Z.-Q. Fang et al., “Deep centers in as-grown and electron-irradiated n-GaN,” in 2000
      International Semiconducting and Insulating Materials Conference. SIMC-XI (Cat.
      No.00CH37046), 35–42, DOI: 10.1109/SIM.2000.939193.
[362] I. Boturchuk et al., Phys. Status Solidi (a), 215, 9, 1700516, 2018, DOI:
      10.1002/pssa.201700516.
[363] A. Y. Polyakov et al., J. Vac. Sci. Technol. B, Nanotechnol. Microelectron. Mater. Process.
      Meas. Phenom., 34, 4, 041216, 2016, DOI: 10.1116/1.4953347.
[364] Z. Zhang et al., Appl. Phys. Lett., 106, 2, 022104, 2015, DOI: 10.1063/1.4905783.
[365] U. Honda et al., Jpn. J. Appl. Phys., 51, 04DF04, 2012, DOI: 10.1143/JJAP.51.04DF04.
[366] Y. Tokuda, ECS Trans., 75, 4, 39–49, 2016, DOI: 10.1149/07504.0039ecst.
[367] H. K. Cho et al., J. Appl. Phys., 94, 3, 1485–1489, 2003, DOI: 10.1063/1.1586981.
[368] H. K. Cho et al., J. Cryst. Growth, 223, 1–2, 38–42, 2001, DOI: 10.1016/S0022-
      0248(00)00982-9.
[369] D. Kindl et al., J. Appl. Phys., 105, 9, 093706, 2009, DOI: 10.1063/1.3122290.
[371] L. Stuchlikova et al., “Investigation of deep energy levels in heterostructures based on GaN
      by DLTS,” in The Eighth International Conference on Advanced Semiconductor Devices
      and Microsystems, 2010, 135–138, DOI: 10.1109/ASDAM.2010.5666319.
[372] P. Hacke et al., J. Appl. Phys., 76, 1, 304–309, 1994, DOI: 10.1063/1.357144.
[373] H. M. Chung et al., Appl. Phys. Lett., 76, 7, 897–899, 2000, DOI: 10.1063/1.125622.
[374] S. Chen et al., J. Appl. Phys., 112, 5, 053513, 2012, DOI: 10.1063/1.4748170.
[375] Y. Nakano, J. Vac. Sci. Technol. A Vacuum, Surfaces, Film., 36, 2, 023001, 2018, DOI:
      10.1116/1.5017867.
[376] G. A. Umana-Membreno et al., Appl. Phys. Lett., 80, 23, 4354–4356, 2002, DOI:
      10.1063/1.1483390.
[377] C. B. Soh et al., J. Phys. Condens. Matter, 16, 34, 6305–6315, 2004, DOI: 10.1088/0953-
      8984/16/34/027.
[378] Y. S. Park et al., Appl. Phys. Lett., 97, 11, 112110, 2010, DOI: 10.1063/1.3491798.
[379] D. K. Johnstonea et al., MRS Proc., 692, H2.7.1, 2001, DOI: 10.1557/PROC-692-H2.7.1.
[380] Z.-Q. Fang et al., Appl. Phys. Lett., 72, 18, 2277–2279, 1998, DOI: 10.1063/1.121274.
[381] K. J. Choi et al., Appl. Phys. Lett., 82, 8, 1233–1235, 2003, DOI: 10.1063/1.1557316.
[382] A. R. Arehart et al., Phys. Status Solidi (c), 5, 6, 1750–1752, 2008, DOI:
      10.1002/pssc.200778622.
[383] H. K. Cho et al., J. Phys. D. Appl. Phys., 41, 15, 155314, 2008, DOI: 10.1088/0022-
      3727/41/15/155314.
[384] A. R. Arehart et al., Appl. Phys. Lett., 96, 24, 242112, 2010, DOI: 10.1063/1.3453660.
[385] O.    Gelhausen    et   al.,   Phys.   Rev.    B,   69,   12,    125210,    2004,      DOI:
      10.1103/PhysRevB.69.125210.
[386] Q. Deng et al., J. Phys. D. Appl. Phys., 44, 34, 345101, 2011, DOI: 10.1088/0022-
      3727/44/34/345101.
[387] M. L. Nakarmi et al., Appl. Phys. Lett., 94, 9, 091903, 2009, DOI: 10.1063/1.3094754.
[388] X. S. Nguyen et al., Phys. Status Solidi (b), 253, 11, 2225–2229, 2016, DOI:
      10.1002/pssb.201600364.
[389] A. Hierro et al., Appl. Phys. Lett., 80, 5, 805–807, 2002, DOI: 10.1063/1.1445274.
[390] A. Y. Polyakov et al., J. Appl. Phys., 95, 10, 5591–5596, 2004, DOI: 10.1063/1.1697616.
[391] E. Calleja et al., Phys. Rev. B, 55, 7, 4689–4694, 1997, DOI: 10.1103/PhysRevB.55.4689.
[392] Z. Zhang et al., Appl. Phys. Lett., 100, 5, 052114, 2012, DOI: 10.1063/1.3682528.
[393] A. Hierro et al., Appl. Phys. Lett., 77, 10, 1499–1501, 2000, DOI: 10.1063/1.1290042.
[394] T. A. Henry et al., Appl. Phys. Lett., 100, 8, 082103, 2012, DOI: 10.1063/1.3687700.
[395] A. Armstrong et al., Opt. Express, 20, S6, A812, 2012, DOI: 10.1364/OE.20.00A812.
[396] Z.-Q. Fang et al., Appl. Phys. Lett., 87, 18, 182115, 2005, DOI: 10.1063/1.2126145.
[397] K. Galiano et al., J. Appl. Phys., 123, 22, 224504, 2018, DOI: 10.1063/1.5022806.
[398] X. D. Chen et al., Appl. Phys. Lett., 82, 21, 3671–3673, 2003, DOI: 10.1063/1.1578167.
[399] A. Sasikumar et al., IEEE Electron Device Lett., 33, 5, 658–660, 2012, DOI:
      10.1109/LED.2012.2188710.
[400] T. Kogiso et al., Jpn. J. Appl. Phys., 58, SC, SCCB36, 2019, DOI: 10.7567/1347-
      4065/ab0408.
[401] A. Y. Polyakov et al., Mater. Sci. Eng. B, 166, 3, 220–224, 2010, DOI:
      10.1016/j.mseb.2009.11.030.
[402] C. D. Wang et al., Appl. Phys. Lett., 72, 10, 1211–1213, 1998, DOI: 10.1063/1.121016.
[403] W. I. Lee et al., Appl. Phys. Lett., 67, 12, 1721–1723, 1995, DOI: 10.1063/1.115028.
[404] M. Caesar et al., “Generation of traps in AlGaN/GaN HEMTs during RF-and DC-stress
      test,” in 2012 IEEE International Reliability Physics Symposium (IRPS), 2012, CD.6.1-
      CD.6.5, DOI: 10.1109/IRPS.2012.6241883.
[405] D. W. Cardwell et al., Appl. Phys. Lett., 102, 19, 193509, 2013, DOI: 10.1063/1.4806980.
[406] M. Silvestri et al., Appl. Phys. Lett., 102, 7, 073501, 2013, DOI: 10.1063/1.4793196.
[407] T. Aggerstam et al., J. Electron. Mater., 36, 12, 1621–1624, 2007, DOI: 10.1007/s11664-
      007-0202-9.
[408] R. Heitz et al., Phys. Rev. B, 55, 7, 4382–4387, 1997, DOI: 10.1103/PhysRevB.55.4382.
[409] A. Sasikumar et al., “Direct correlation between specific trap formation and electric stress-
      induced degradation in MBE-grown AlGaN/GaN HEMTs,” in 2012 IEEE International
      Reliability     Physics      Symposium       (IRPS),         2012,      2C.3.1-2C.3.6,    DOI:
      10.1109/IRPS.2012.6241780.
[410] Z. Zhang et al., Appl. Phys. Lett., 101, 15, 152104, 2012, DOI: 10.1063/1.4759037.
[411] I.-H. Lee et al., Appl. Phys. Lett., 110, 19, 192107, 2017, DOI: 10.1063/1.4983556.
[413] M. Huber et al., Appl. Phys. Lett., 107, 3, 032106, 2015, DOI: 10.1063/1.4927405.
[414] W. Götz et al., Appl. Phys. Lett., 68, 22, 3144–3146, 1996, DOI: 10.1063/1.115805.
[415] M.    Kumar     et   al.,   Mater.   Res.   Bull.,     47,    6,     1306–1309,   2012,   DOI:
      10.1016/j.materresbull.2012.03.016.
[416] Y. Irokawa et al., J. Appl. Phys., 97, 8, 083505, 2005, DOI: 10.1063/1.1863458.
[417] T. Tanaka et al., Appl. Phys. Lett., 65, 5, 593–594, 1994, DOI: 10.1063/1.112309.
[418] H. Nakayama et al., Jpn. J. Appl. Phys., 35, Part 2, No. 3A, L282–L284, 1996, DOI:
      10.1143/JJAP.35.L282.
[419] M. Horita et al., Appl. Phys. Express, 13, 7, 071007, 2020, DOI: 10.35848/1882-
      0786/ab9e7c.
[420] K. Sharma et al., “Impact of the Location of Iron Buffer Doping on Trap Signatures in GaN
      HEMTs,” in 2020 International Workshop on Integrated Nonlinear Microwave and
      Millimetre-Wave          Circuits        (INMMiC),          2020,         1–3,      DOI:
      10.1109/INMMiC46721.2020.9160114.
[421] J. L. Lyons et al., Phys. Rev. B, 89, 3, 035204, 2014, DOI: 10.1103/PhysRevB.89.035204.
[422] E. Kohn et al., IEEE Trans. Microw. Theory Tech., 51, 2, 634–642, 2003, DOI:
      10.1109/TMTT.2002.807687.
[423] W. S. Tan et al., IEEE Electron Device Lett., 27, 1, 1–3, 2006, DOI:
      10.1109/LED.2005.860383.
[424] M. Meneghini et al., IEEE Trans. Power Electron., 29, 5, 2199–2207, 2014, DOI:
      10.1109/TPEL.2013.2271977.
[425] M. Neuburger et al., IEEE Electron Device Lett., 25, 5, 256–258, 2004, DOI:
      10.1109/LED.2004.827283.
[426] G. Meneghesso et al., Semicond. Sci. Technol., 28, 7, 074021, 2013, DOI: 10.1088/0268-
      1242/28/7/074021.
[427] B. M. Green et al., IEEE Electron Device Lett., 21, 6, 268–270, 2000, DOI:
      10.1109/55.843146.
[429] N.-Q. Zhang et al., IEEE Electron Device Lett., 21, 9, 421–423, 2000, DOI:
      10.1109/55.863096.
[430] A. Chini et al., Electron. Lett., 40, 1, 73, 2004, DOI: 10.1049/el:20040017.
[431] M. Meneghini et al., Appl. Phys. Lett., 100, 23, 233508, 2012, DOI: 10.1063/1.4723848.
[432] D.   Jin   et   al.,   Microelectron.   Reliab.,   52,   12,   2875–2879,   2012,    DOI:
      10.1016/j.microrel.2012.08.023.
[433] I. Hwang et al., IEEE Electron Device Lett., 34, 12, 1494–1496, 2013, DOI:
      10.1109/LED.2013.2286173.
[434] I. Rossetto et al., IEEE Trans. Electron Devices, 64, 9, 3734–3739, 2017, DOI:
      10.1109/TED.2017.2728785.
[435] M. Higashiwaki et al., J. Appl. Phys., 108, 6, 063719, 2010, DOI: 10.1063/1.3481412.
[436] M. Matys et al., J. Appl. Phys., 120, 22, 225305, 2016, DOI: 10.1063/1.4971409.
[437] E. Ber et al., IEEE Trans. Electron Devices, 66, 5, 2100–2105, 2019, DOI:
      10.1109/TED.2019.2901869.
[438] B. Bakeroot et al., J. Appl. Phys., 116, 13, 134506, 2014, DOI: 10.1063/1.4896900.
[439] M. Faqir et al., Microelectron. Reliab., 50, 9–11, 1520–1522, 2010, DOI:
      10.1016/j.microrel.2010.07.020.
[440] M. Grupen, IEEE Trans. Electron Devices, 66, 9, 3777–3783, 2019, DOI:
      10.1109/TED.2019.2928536.
[441] M. J. Uren et al., IEEE Trans. Electron Devices, 53, 2, 395–398, 2006, DOI:
      10.1109/TED.2005.862702.
[442] E. Bahat-Treidel et al., IEEE Trans. Electron Devices, 57, 11, 3050–3058, 2010, DOI:
      10.1109/TED.2010.2069566.
[443] M. J. Uren et al., IEEE Trans. Electron Devices, 59, 12, 3327–3333, 2012, DOI:
      10.1109/TED.2012.2216535.
[444] C. Zhou et al., IEEE Electron Device Lett., 33, 8, 1132–1134, 2012, DOI:
      10.1109/LED.2012.2200874.
[445] M. Wang et al., IEEE Electron Device Lett., 35, 11, 1094–1096, 2014, DOI:
      10.1109/LED.2014.2356720.
[446] A. Guo et al., “Negative-bias temperature instability of GaN MOSFETs,” in 2016 IEEE
      International Reliability Physics Symposium (IRPS), 2016, 4A-1-1-4A-1–6, DOI:
      10.1109/IRPS.2016.7574526.
[447] D. Jin et al., “Total current collapse in high-voltage GaN MIS-HEMTs induced by Zener
      trapping,” in 2013 IEEE International Electron Devices Meeting, 2013, 6.2.1-6.2.4, DOI:
      10.1109/IEDM.2013.6724572.
[448] M. J. Uren et al., IEEE Trans. Electron Devices, 64, 7, 2826–2834, 2017, DOI:
      10.1109/TED.2017.2706090.
[449] N. Zagni et al., “Trap Dynamics Model Explaining the R ON Stress/Recovery Behavior in
      Carbon-Doped Power AlGaN/GaN MOS-HEMTs,” in 2020 IEEE International Reliability
      Physics Symposium (IRPS), 2020, 1–5, DOI: 10.1109/IRPS45951.2020.9128816.
[450] J. Neugebauer et al., “Native defects and impurities in GaN,” 1996, 25–44.
[451] P. B. Klein et al., Appl. Phys. Lett., 79, 21, 3527–3529, 2001, DOI: 10.1063/1.1418452.
[452] Y.-F. Wu et al., IEEE Electron Device Lett., 25, 3, 117–119, 2004, DOI:
      10.1109/LED.2003.822667.
[453] A. Armstrong et al., Appl. Phys. Lett., 88, 8, 082114, 2006, DOI: 10.1063/1.2179375.
[454] X. Li et al., J. Vac. Sci. Technol. B, Nanotechnol. Microelectron. Mater. Process. Meas.
      Phenom., 33, 2, 021208, 2015, DOI: 10.1116/1.4914316.
[455] A. Armstrong et al., J. Appl. Phys., 98, 5, 053704, 2005, DOI: 10.1063/1.2005379.
[457] J. L. Lyons et al., Appl. Phys. Lett., 97, 15, 152108, 2010, DOI: 10.1063/1.3492841.
[458] M. Matsubara et al., J. Appl. Phys., 121, 19, 195701, 2017, DOI: 10.1063/1.4983452.
[459] G. Meneghesso et al., “Threshold voltage instabilities in D-mode GaN HEMTs for power
      switching applications,” in 2014 IEEE International Reliability Physics Symposium, 2014,
      6C.2.1-6C.2.5, DOI: 10.1109/IRPS.2014.6861109.
[460] A. Chini et al., IEEE Trans. Electron Devices, 63, 9, 3473–3478, 2016, DOI:
      10.1109/TED.2016.2593791.
[461] V. Joshi et al., IEEE Trans. Electron Devices, 66, 1, 561–569, 2019, DOI:
      10.1109/TED.2018.2878770.
[463] N.   Zagni    et    al.,    Phys.   Status   Solidi   (a),   217,   7,   1900762,   2020,   DOI:
      10.1002/pssa.201900762.
[464] N. Zagni et al., J. Comput. Electron., 19, 4, 1555–1563, 2020, DOI: 10.1007/s10825-020-
      01573-8.
[465] A. Fariza et al., Appl. Phys. Lett., 109, 21, 212102, 2016, DOI: 10.1063/1.4968823.
[466] I. Chatterjee et al., “Impact of buffer charge on the reliability of carbon doped AlGaN/GaN-
      on-Si HEMTs,” in 2016 IEEE International Reliability Physics Symposium (IRPS), 2016,
      4A-4-1-4A-4–5, DOI: 10.1109/IRPS.2016.7574529.
[467] B. Rackauskas et al., IEEE Trans. Electron Devices, 65, 5, 1838–1842, 2018, DOI:
      10.1109/TED.2018.2813542.
[468] M. Singh et al., IEEE Trans. Electron Devices, 65, 9, 3746–3753, 2018, DOI:
      10.1109/TED.2018.2860902.
[469] C. Koller et al., IEEE Trans. Electron Devices, 65, 12, 5314–5321, 2018, DOI:
      10.1109/TED.2018.2872552.
[470] C. Poblenz et al., J. Vac. Sci. Technol. B Microelectron. Nanom. Struct., 22, 3, 1145, 2004,
      DOI: 10.1116/1.1752907.
[471] A. Nardo et al., Appl. Phys. Express, 13, 7, 074003, 2020, DOI: 10.35848/1882-
      0786/ab9623.
[472] J. He et al., IEEE Trans. Electron Devices, 65, 8, 3185–3191, 2018, DOI:
      10.1109/TED.2018.2850042.
[473] F. Sang et al., Jpn. J. Appl. Phys., 54, 4, 044101, 2015, DOI: 10.7567/JJAP.54.044101.
[474] M.    Ruzzarin   et    al.,     Microelectron.     Reliab.,   100–101,     113488,   2019,   DOI:
      10.1016/j.microrel.2019.113488.
[475] A. G. Viey et al., IEEE Trans. Electron Devices, 68, 4, 2017–2024, 2021, DOI:
      10.1109/TED.2021.3050127.
[476] G. L. Bilbro et al., Electron. Lett., 42, 24, 1425, 2006, DOI: 10.1049/el:20062113.
[477] Hou-Yu-Wang et al., “Effects of the Fe-doped GaN buffer in AlGaN/GaN HEMTs on SiC
      substrate,” in 2015 IEEE International Conference on Electron Devices and Solid-State
      Circuits (EDSSC), 2015, 645–648, DOI: 10.1109/EDSSC.2015.7285198.
[478] F. Iucolano et al., “Correlation between dynamic Rdsou transients and Carbon related buffer
      traps in AlGaN/GaN HEMTs,” in 2016 IEEE International Reliability Physics Symposium
      (IRPS), 2016, CD-2-1-CD-2-4, DOI: 10.1109/IRPS.2016.7574586.
[479] K. Tanaka et al., Jpn. J. Appl. Phys., 52, 4S, 04CF07, 2013, DOI: 10.7567/JJAP.52.04CF07.
[480] D. Bisi et al., Phys. Status Solidi (a), 212, 5, 1122–1129, 2015, DOI:
      10.1002/pssa.201431744.
[481] G. Meneghesso et al., Semicond. Sci. Technol., 31, 9, 093004, 2016, DOI: 10.1088/0268-
      1242/31/9/093004.
[482] N. Zagni et al., IEEE Trans. Electron Devices, 68, 2, 697–703, 2021, DOI:
      10.1109/TED.2020.3045683.
[484] P. Moens et al., “Impact of buffer leakage on intrinsic reliability of 650V AlGaN/GaN
      HEMTs,” in 2015 IEEE International Electron Devices Meeting (IEDM), 2015, 35.2.1-
      35.2.4, DOI: 10.1109/IEDM.2015.7409831.
[486] M. J. Uren et al., IEEE Electron Device Lett., 35, 3, 327–329, 2014, DOI:
      10.1109/LED.2013.2297626.
[487] M. Meneghini et al., Mater. Sci. Semicond. Process., 78, 118–126, 2018, DOI:
      10.1016/j.mssp.2017.10.009.
[488] M. Van Hove et al., IEEE Electron Device Lett., 33, 5, 667–669, 2012, DOI:
      10.1109/LED.2012.2188016.
[489] J. A. del Alamo et al., IEEE Trans. Electron Devices, 66, 11, 4578–4590, 2019, DOI:
      10.1109/TED.2019.2931718.
[490] M. Hua et al., IEEE Electron Device Lett., 39, 3, 413–416, 2018, DOI:
      10.1109/LED.2018.2791664.
[492] M. Hua et al., IEEE Trans. Electron Devices, 65, 9, 3831–3838, 2018, DOI:
      10.1109/TED.2018.2856998.
[493] M. Hua et al., “Reverse-bias stability and reliability of hole-barrier-free E-mode LPCVD-
      SiN<inf>x</inf>/GaN MIS-FETs,” in 2017 IEEE International Electron Devices Meeting
      (IEDM), 2017, 33.2.1-33.2.4, DOI: 10.1109/IEDM.2017.8268489.
[494] A. Guo et al., “Positive-bias temperature instability (PBTI) of GaN MOSFETs,” in 2015
      IEEE International Reliability Physics Symposium, 2015, 6C.5.1-6C.5.7, DOI:
      10.1109/IRPS.2015.7112770.
[495] T.-L. Wu et al., IEEE Trans. Electron Devices, 63, 5, 1853–1860, 2016, DOI:
      10.1109/TED.2016.2539341.
[496] L. Sayadi et al., IEEE Trans. Electron Devices, 65, 6, 2454–2460, 2018, DOI:
      10.1109/TED.2018.2828702.
[497] I. Rossetto et al., Microelectron. Reliab., 55, 9–10, 1692–1696, 2015, DOI:
      10.1016/j.microrel.2015.06.130.
[498] J. K. Kaushik et al., IEEE Trans. Electron Devices, 60, 10, 3351–3357, 2013, DOI:
      10.1109/TED.2013.2279158.
[499] P. Altuntas et al., “On the correlation between kink effect and effective mobility in
      InAlN/GaN HEMTs,” in 2014 9th European Microwave Integrated Circuit Conference,
      2014, 88–91, DOI: 10.1109/EuMIC.2014.6997798.
[500] S. D. Nsele et al., IEEE Trans. Electron Devices, 60, 4, 1372–1378, 2013, DOI:
      10.1109/TED.2013.2248158.
[501] X.-H. Ma et al., Chinese Phys. B, 23, 2, 027302, 2014, DOI: 10.1088/1674-
      1056/23/2/027302.
[503] H. Hirshy et al., IEEE Trans. Electron Devices, 65, 12, 5307–5313, 2018, DOI:
      10.1109/TED.2018.2872513.
[504] A. Chini et al., Microelectron. Reliab., 54, 9–10, 2222–2226, 2014, DOI:
      10.1016/j.microrel.2014.07.085.
[505] O. Axelsson et al., IEEE Trans. Electron Devices, 63, 1, 326–332, 2016, DOI:
      10.1109/TED.2015.2499313.
[507] C. T. Sah et al., Solid. State. Electron., 13, 6, 759–788, 1970, DOI: 10.1016/0038-
      1101(70)90064-X.
[509] A. M. White et al., J. Appl. Phys., 47, 7, 3230–3239, 1976, DOI: 10.1063/1.323120.
[510] A. Chantre et al., Phys. Rev. B, 23, 10, 5335–5359, 1981, DOI: 10.1103/PhysRevB.23.5335.
[511] A. Hierro et al., Appl. Phys. Lett., 76, 21, 3064–3066, 2000, DOI: 10.1063/1.126580.
[512] Z. Zhang et al., J. Appl. Phys., 118, 15, 155701, 2015, DOI: 10.1063/1.4933174.
[513] O. Mitrofanov et al., Superlattices Microstruct., 34, 1–2, 33–53, 2003, DOI:
      10.1016/j.spmi.2003.12.002.
[514] T. Mizutani et al., Phys. Status Solidi (a), 200, 1, 195–198, 2003, DOI:
      10.1002/pssa.200303464.
[515] T. Okino et al., IEEE Electron Device Lett., 25, 8, 523–525, 2004, DOI:
      10.1109/LED.2004.832788.
[516] A. R. Arehart et al., “Spatially-discriminating trap characterization methods for HEMTs and
      their application to RF-stressed AlGaN/GaN HEMTs,” in 2010 International Electron
      Devices Meeting, 2010, 20.1.1-20.1.4, DOI: 10.1109/IEDM.2010.5703396.
[517] D. Bisi et al., IEEE Trans. Electron Devices, 60, 10, 3166–3175, 2013, DOI:
      10.1109/TED.2013.2279021.
[518] N. Modolo et al., Semicond. Sci. Technol., 36, 1, 014001, 2021, DOI: 10.1088/1361-
      6641/abc456.
[519] D. Bisi et al., “High-voltage double-pulsed measurement system for GaN-based power
      HEMTs,” in 2014 IEEE International Reliability Physics Symposium, 2014, CD.11.1-
      CD.11.4, DOI: 10.1109/IRPS.2014.6861130.
[520] J. Joh et al., IEEE Trans. Electron Devices, 58, 1, 132–140, 2011, DOI:
      10.1109/TED.2010.2087339.
[521] M. Tapajna et al., IEEE Electron Device Lett., 31, 7, 662–664, 2010, DOI:
      10.1109/LED.2010.2047092.
[522] G. Meneghesso et al., IEEE Trans. Electron Devices, 51, 10, 1554–1561, 2004, DOI:
      10.1109/TED.2004.835025.
[523] A. Chini et al., “Experimental and numerical correlation between current-collapse and fe-
      doping profiles in GaN HEMTs,” in 2012 IEEE International Reliability Physics
      Symposium (IRPS), 2012, CD.2.1-CD.2.4, DOI: 10.1109/IRPS.2012.6241881.
[524] A. Chini et al., IEEE Trans. Electron Devices, 60, 10, 3176–3182, 2013, DOI:
      10.1109/TED.2013.2278290.
[525] O. Mitrofanov et al., J. Appl. Phys., 95, 11, 6414–6419, 2004, DOI: 10.1063/1.1719264.
[526] A. Chini et al., Mater. Sci. Semicond. Process., 78, 127–131, 2018, DOI:
      10.1016/j.mssp.2017.10.029.
[528] C. N. Berglund, IEEE Trans. Electron Devices, ED-13, 10, 701–705, 1966, DOI:
      10.1109/T-ED.1966.15827.
[529] H. Mazari et al., J. New Technol. Mater., 8, 1, 97–101, 2018, DOI: 10.12816/0048929.
[530] E. H. Nicollian et al., Bell Syst. Tech. J., 46, 6, 1055–1133, 1967, DOI: 10.1002/j.1538-
      7305.1967.tb01727.x.
[531] W. A. Hill et al., Solid. State. Electron., 23, 9, 987–993, 1980, DOI: 10.1016/0038-
      1101(80)90064-7.
[533] Z.-F. Zhu et al., Chinese Phys. Lett., 34, 9, 097301, 2017, DOI: 10.1088/0256-
      307X/34/9/097301.
[536] T. S. Lay et al., Electron. Lett., 37, 9, 595, 2001, DOI: 10.1049/el:20010403.
[537] S.-B. Bae et al., Microelectron. Eng., 109, 10–12, 2013, DOI: 10.1016/j.mee.2013.03.108.
[538] M. Cui et al., “Characterization of Transient Threshold Voltage Shifts in Enhancement- and
      Depletion-mode AlGaN/GaN Metal-Insulator-Semiconductor (MIS)-HEMTs,” in 2018
      IEEE International Conference on Electron Devices and Solid State Circuits (EDSSC),
      2018, 1–2, DOI: 10.1109/EDSSC.2018.8487160.
[539] M. Miczek et al., J. Appl. Phys., 103, 10, 104510, 2008, DOI: 10.1063/1.2924334.
[540] M. Capriotti et al., J. Appl. Phys., 117, 2, 024506, 2015, DOI: 10.1063/1.4905945.
[541] B. L. Swenson et al., J. Appl. Phys., 106, 6, 064902, 2009, DOI: 10.1063/1.3224852.
[543] M. A. Reshchikov et al., J. Appl. Phys., 97, 6, 061301, 2005, DOI: 10.1063/1.1868059.
[544] T. Ogino et al., Jpn. J. Appl. Phys., 19, 12, 2395–2405, 1980, DOI: 10.1143/JJAP.19.2395.
[545] M.    Julkarnain        et          al.,     Opt.    Mater.        (Amst).,     60,   481–486,        2016,   DOI:
      10.1016/j.optmat.2016.09.003.
[548] R. Degraeve et al., Microelectron. Reliab., 39, 10, 1445–1460, 1999, DOI: 10.1016/S0026-
      2714(99)00051-7.
[551] T.   Wu    et   al.,         2015          IEEE      Int.   Reliab.     Phys.     Symp.,    4–9,      2015,   DOI:
      10.1109/IRPS.2015.7112769.
[553] K. Mukherjee et al., IEEE Int. Reliab. Phys. Symp. Proc., 2020-April, 2–6, 2020, DOI:
      10.1109/IRPS45951.2020.9129098.
[554] M. Ruzzarin et al., Microelectron. Reliab., 88–90, June, 620–626, 2018, DOI:
      10.1016/j.microrel.2018.06.044.
[555] M. Ťapajna et al., Appl. Phys. Lett., 107, 19, 193506, 2015, DOI: 10.1063/1.4935223.
[556] I. Rossetto et al., IEEE Trans. Electron Devices, 63, 6, 2334–2339, 2016, DOI:
      10.1109/TED.2016.2553721.
[557] T.-L. Wu et al., IEEE Electron Device Lett., 36, 10, 1001–1003, 2015, DOI:
      10.1109/LED.2015.2465137.
[562] F. Masin et al., Appl. Phys. Lett., 115, 5, 052103, 2019, DOI: 10.1063/1.5109301.
[563] J. He et al., IEEE Trans. Electron Devices, 66, 8, 3453–3458, 2019, DOI:
      10.1109/TED.2019.2924675.
[564] I. C. Kizilyalli et al., Microelectron. Reliab., 55, 9–10, 1654–1661, 2015, DOI:
      10.1016/j.microrel.2015.07.012.
[566] K. Orita et al., IEEE J. Quantum Electron., 48, 9, 1169–1176, 2012, DOI:
      10.1109/JQE.2012.2203795.
[567] E. Zanoni et al., “Reliability of Gallium Nitride microwave transistors,” in 2016 21st
      International Conference on Microwave, Radar and Wireless Communications (MIKON),
      2016, 1–6, DOI: 10.1109/MIKON.2016.7492013.
[570] D. Bisi et al., IEEE Electron Device Lett., 36, 10, 1011–1014, 2015, DOI:
      10.1109/LED.2015.2474116.
[571] M.    Caesar    et    al.,    IEEE    Int.    Reliab.      Phys.   Symp.,   1–5,   2012,   DOI:
      10.1109/IRPS.2012.6241883.
[574] M. Ruzzarin et al., “Degradation of vertical GaN FETs under gate and drain stress,” in 2018
      IEEE International Reliability Physics Symposium (IRPS), 2018, 4B.1-1-4B.1-5, DOI:
      10.1109/IRPS.2018.8353579.
[575] M. Meneghini et al., IEEE Trans. Electron Devices, 62, 8, 2549–2554, 2015, DOI:
      10.1109/TED.2015.2446032.
[576] I. Rossetto et al., IEEE Trans. Electron Devices, 65, 4, 1303–1307, 2018, DOI:
      10.1109/TED.2018.2802449.
[577] I. L. Guy et al., Appl. Phys. Lett., 75, 26, 4133–4135, 1999, DOI: 10.1063/1.125560.
[578] M. A. der Maur et al., IEEE Trans. Electron Devices, 60, 10, 3142–3148, 2013, DOI:
      10.1109/TED.2013.2267547.
[579] M. Power et al., “Reverse-biased induced mechanical stress in AlGaN/GaN power diodes,”
      in 2016 28th International Symposium on Power Semiconductor Devices and ICs (ISPSD),
      2016, 103–106, DOI: 10.1109/ISPSD.2016.7520788.
[580] J. Joh et al., “Mechanisms for Electrical Degradation of GaN High-Electron Mobility
      Transistors,” in 2006 International Electron Devices Meeting, 2006, 29, 4, 1–4, DOI:
      10.1109/IEDM.2006.346799.
[581] J. Joh et al., “Impact of electrical degradation on trapping characteristics of GaN high
      electron mobility transistors,” in 2008 IEEE International Electron Devices Meeting, 2008,
      1–4, DOI: 10.1109/IEDM.2008.4796725.
[582] J. Joh et al., IEEE Electron Device Lett., 29, 4, 287–289, 2008, DOI:
      10.1109/LED.2008.917815.
[583] C.-Y. Chang et al., IEEE Trans. Device Mater. Reliab., 11, 1, 187–193, 2011, DOI:
      10.1109/TDMR.2010.2103314.
[584] P. Makaram et al., Appl. Phys. Lett., 96, 23, 233509, 2010, DOI: 10.1063/1.3446869.
[585] F. Gao et al., Appl. Phys. Lett., 99, 22, 223506, 2011, DOI: 10.1063/1.3665065.
[586] F. Gao et al., IEEE Trans. Electron Devices, 61, 2, 437–444, 2014, DOI:
      10.1109/TED.2013.2293114.
[587] M. M. Bajo et al., Appl. Phys. Lett., 104, 22, 223506, 2014, DOI: 10.1063/1.4881637.
[589] P. Moens et al., Tech. Dig. - Int. Electron Devices Meet. IEDM, 2016-Febru, 35.2.1-35.2.4,
      2015, DOI: 10.1109/IEDM.2015.7409831.
[590] C. De Santi et al., IEEE Electron Device Lett., 37, 5, 611–614, 2016, DOI:
      10.1109/LED.2016.2543805.
[591] M. Borga et al., IEEE Trans. Electron Devices, 64, 9, 3616–3621, 2017, DOI:
      10.1109/TED.2017.2726440.
[592] M. Borga et al., Microelectron. Reliab., 100–101, September, 113461, 2019, DOI:
      10.1016/j.microrel.2019.113461.
[593] A. Tajalli et al., Micromachines, 11, 1, 2020, DOI: 10.3390/mi11010101.
[594] M. Borga et al., IEEE Trans. Electron Devices, 67, 2, 595–599, 2020, DOI:
      10.1109/TED.2020.2964060.
[596] C. De Santi et al., IEEE Electron Device Lett., 41, 9, 1300–1303, 2020, DOI:
      10.1109/LED.2020.3009649.
[597] E. Fabris et al., IEEE Trans. Electron Devices, 66, 11, 4597–4603, 2019, DOI:
      10.1109/TED.2019.2943014.
[599] A. Sasikumar et al., IEEE Int. Reliab. Phys. Symp. Proc., 1–6, 2012, DOI:
      10.1109/IRPS.2012.6241780.
[600] A. Minetto et al., IEEE Trans. Electron Devices, 67, 11, 4602–4605, 2020, DOI:
      10.1109/TED.2020.3025983.
[601] F. Yang et al., IEEE Trans. Electron Devices, 67, 3, 869–874, 2020, DOI:
      10.1109/TED.2020.2968212.
[602] J. Joh et al., “Current collapse in GaN heterojunction field effect transistors for high-voltage
      switching applications,” in 2014 IEEE International Reliability Physics Symposium, 2014,
      6C.5.1-6C.5.4, DOI: 10.1109/IRPS.2014.6861112.
[604] A. Barbato et al., IET Power Electron., 13, 11, 2390–2397, 2020, DOI: 10.1049/iet-
      pel.2019.1455.
[605] N. Modolo et al., IEEE Electron Device Lett., 42, 5, 673–676, 2021, DOI:
      10.1109/LED.2021.3067796.
[606] M. Meneghini et al., IEEE Electron Device Lett., 33, 3, 375–377, 2012, DOI:
      10.1109/LED.2011.2181815.
[607] J. Kuzmı́k et al., Solid. State. Electron., 48, 2, 271–276, 2004, DOI: 10.1016/S0038-
      1101(03)00295-8.
[608] Shih-Hung Chen et al., IEEE Trans. Device Mater. Reliab., 12, 4, 589–598, 2012, DOI:
      10.1109/TDMR.2012.2217746.
[609] B. Shankar et al., “Unique ESD behavior and failure modes of AlGaN/GaN HEMTs,” in
      2016 IEEE International Reliability Physics Symposium (IRPS), 2016, EL-7-1-EL-7-5,
      DOI: 10.1109/IRPS.2016.7574608.
[610] I. Rossetto et al., IEEE Trans. Electron Devices, 62, 9, 2830–2836, 2015, DOI:
      10.1109/TED.2015.2463713.
[611] C. De Santi et al., IET Power Electron., 11, 4, 668–674, 2018, DOI: 10.1049/iet-
      pel.2017.0403.
[613] S. J. Pearton et al., ECS J. Solid State Sci. Technol., 5, 2, Q35–Q60, 2016, DOI:
      10.1149/2.0251602jss.
[614] A. Y. Polyakov et al., Mater. Sci. Eng. R Reports, 94, 1–56, 2015, DOI:
      10.1016/j.mser.2015.05.001.
[615] A. Stockman et al., IEEE Trans. Electron Devices, 66, 1, 372–377, 2019, DOI:
      10.1109/TED.2018.2881325.
[616] X.   Hu    et    al.,   IEEE   Trans.   Nucl.   Sci.,   51,   2,   293–297,   2004,   DOI:
      10.1109/TNS.2004.825077.
[617] L. Liu et al., J. Vac. Sci. Technol. B, Nanotechnol. Microelectron. Mater. Process. Meas.
      Phenom., 31, 2, 022201, 2013, DOI: 10.1116/1.4788904.
[618] B. D. White et al., IEEE Trans. Nucl. Sci., 50, 6, 1934–1941, 2003, DOI:
      10.1109/TNS.2003.821827.
[619] Xinwen Hu et al., IEEE Trans. Nucl. Sci., 50, 6, 1791–1796, 2003, DOI:
      10.1109/TNS.2003.820792.
[620] A. P. Karmarkar et al., IEEE Trans. Nucl. Sci., 51, 6, 3801–3806, 2004, DOI:
      10.1109/TNS.2004.839199.
[621] B. Luo et al., Appl. Phys. Lett., 79, 14, 2196–2198, 2001, DOI: 10.1063/1.1408606.
[622] B. Luo et al., Solid. State. Electron., 47, 6, 1015–1020, 2003, DOI: 10.1016/S0038-
      1101(02)00468-9.
[623] M. P. King et al., IEEE Trans. Nucl. Sci., 62, 6, 2912–2918, 2015, DOI:
      10.1109/TNS.2015.2480071.
[624] J.   Chen et    al.,   IEEE Trans.    Nucl.   Sci., 60, 6, 4080–4086, 2013, DOI:
      10.1109/TNS.2013.2281771.
[625] B. Luo et al., J. Electron. Mater., 31, 5, 437–441, 2002, DOI: 10.1007/s11664-002-0097-4.
[626] A. Goyal et al., J. Appl. Phys., 117, 22, 225702, 2015, DOI: 10.1063/1.4922286.
[627] B. Luo et al., Appl. Phys. Lett., 80, 4, 604–606, 2002, DOI: 10.1063/1.1445809.
[628] A. Yadav et al., Radiat. Eff. Defects Solids, 170, 5, 377–385, 2015, DOI:
      10.1080/10420150.2015.1010170.
[629] C. Schwarz et al., Appl. Phys. Lett., 102, 6, 062102, 2013, DOI: 10.1063/1.4792240.
[630] S. A. Vitusevich et al., Phys. Status Solidi (a), 195, 1, 101–105, 2003, DOI:
      10.1002/pssa.200306264.
[631] B. D. Weaver et al., IEEE Trans. Nucl. Sci., 59, 6, 3077–3080, 2012, DOI:
      10.1109/TNS.2012.2224371.