2011 Balaz PHD
2011 Balaz PHD
2011 Balaz PHD
http://theses.gla.ac.uk/2676/
Copyright and moral rights for this thesis are retained by the author
The content must not be changed in any way or sold commercially in any
format or medium without the formal permission of the Author
Daniel Balaz
It is widely believed that the current collapse is caused by a virtual gate, i.e. electrons
leaked to the surface of the device. We have found a charge distribution that reproduced
the I-V measurement that shows current collapse, hence validating the concept of the
virtual gate. While it was previously shown that the virtual gate has a similar impact on the
I-V curve as is observed during the current collapse, we believe that this is for the first time
that a wide range of gate and drain voltages was calibrated.
High gate/drain voltage leading to permanent degradation was also investigated. The
hypothesis that stress induced defects and dislocations might be responsible for the
degradation was tested but not fully confirmed.
Finally, the leakage of electrons thought to be responsible for formation of the virtual gate
and the current collapse due to the Poole-Frenkel emission, is simulated in order to explain
the surface charge distribution responsible for the current collapse and deduced in
Chapter 5.
Acknowledgements
The three most important things in completing a strenuous journey are perseverance and
luck, the latter of which oftentimes manifests itself in the people the traveller meets. In my
case, the following people helped to make this endeavour successful. First of all, I would
like to thank Prof. Asen Asenov, my supervisor, for his guidance and advice during my
study, for sharing his knowledge and very prompt responding to any communication. Also,
for having trust in me completing this thesis even in times of my doubts. At least, he gave
that impression. I am thankful to Dr Karol Kalna, my second supervisor, for bringing the
opportunity to study here to my attention, helping me to be accepted and then for being
ready and willing to discuss the questions arising from my research and for reading my
output.
I wish to thank to Prof. Martin Kuball (University of Bristol) for initiating this project and
for discussions at our meetings and to Prof. Michael J. Uren (QinetiQ, now at the
University of Bristol) for providing the measurements used in this work and for invaluable
discussions and advice on the operation of HEMT devices. Also, thanks to Dr Andrei
Sarua (University of Bristol) for the interchange of ideas in emails.
I am also thankful to other members of the Device Modelling Group, of which I would like
to mention: Dr Ewan Towie for being the first Scotsman to engage in discussions with me
outside the scope of our study, Iain Moore for having many interesting and fun
conversations which undoubtedly delayed the finishing of my thesis, Dr Stanislav Markov
for offering his help and advice frequently and with enthusiasm, and Dr Urban Kovac, the
group‟s mathematical expert, for refreshing ideas on solving some problems.
Special thanks go to my parents for their help, support and encouragement. Finally, many
thanks to Izabela for moving to cold, dark and rainy Scotland for a few years, which is a
bigger challenge for her than for me, for insisting on me finishing this thesis, for delicious
dinners but, most importantly, for giving birth to our lovely daughter Sophie Isabel.
List of Conferences and Publications
D. Balaz, K. Kalna, M. Kuball, M. J. Uren, and A. Asenov, "Impact of the field induced
polarization space-charge on the characteristics of AlGaN/GaN HEMT: Self-consistent
simulation study," in IWN. International Workshop on Nitrides, 2008.
D. Balaz, K. Kalna, M. Kuball, M. J. Uren, and A. Asenov, "Impact of the field induced
polarization space-charge on the characteristics of AlGaN/GaN HEMT: Self-consistent
simulation study," Physica Status Solidi (c), vol. 6, no. S2, pp. S1007-S1011, 2009.
1 INTRODUCTION ............................................................................................................... 1
2.4.3 Substrates............................................................................................... 35
2.4.3.1 SiC ......................................................................................... 35
2.4.3.3 Si ........................................................................................... 36
2.5.1 Self-Heating........................................................................................... 37
2.5.2 Current Collapse and Degradation ........................................................ 38
2.6 SUMMARY ............................................................................................................ 39
6.5.3 The impact of the Time Step on the Surface Charge and Leakage
Current 136
6.6 SUMMARY .......................................................................................................... 138
List of Figures
Figure 1.1: Summary of failure mechanisms in AlGaN/GaN HEMTs. Taken from [17]. .... 3
Figure 2.1: III-Nitrides crystallize in cubic zinc-blende (left) and hexagonal wurtzite (right)
structure. They both lack the centre of symmetry, so they show piezoelectricity. The
wurtzite, which is a lower symmetry crystal, possesses also the spontaneous
polarization. The lattice parameters a, c and u are shown for the wurtzite structure. The
arrows on the zinc-blende structure depict the set of <111> directions. Adapted from
[37]. ................................................................................................................................... 8
Figure 2.2: Ga- and N- face of GaN. AlN and InN show similar structure. The c-axis is the
polar axis in wurtzite crystals. The group III (Al, Ga, In) and N atoms along this bond
form the dipoles, which are the basis for the macroscopic polarization. Taken from [41]9
Figure 2.3: The relationships between electrical and mechanical properties of a crystal. The
names of the variables (ovals) and the properties (squares) and their corresponding
symbols that will be used in this work are shown. The rank of the tensors representing
the corresponding variables and properties can be deduced from the number of their
indices. Adapted from [46]. ............................................................................................. 14
Figure 2.4: Spontaneous polarization of III-N binary and ternary alloys. The thin lines
show linear Vegard-like interpolation, thick lines show the approximation to second
order in disordered ternary nitride alloys, calculated using the parabolic model (2.1) and
the previously mentioned bowing parameters [53]. The higher is the mismatch between
LIST OF FIGURES xi
the lattice constants of the parent binaries, the higher is the spontaneous polarization
bowing. ............................................................................................................................ 17
Figure 2.5: A heterojunction of two different semiconductors, which have different band-
gap energies Eg and affinities χ. The difference of the affinities determines the
conduction band discontinuity ΔEc and that together with the band-gap energy
difference determines the valence band discontinuity ΔEv. In a heterostructure, the
electrons and holes see a different barrier height. Doping of either of the
semiconductors would result in band bending. However, in III-N based devices, the
channel electrons are provided by the polarization difference of the two semiconductors,
which is sometimes referred to as polarization doping. Hence there is no need for
doping. ............................................................................................................................. 19
Figure 2.6: The impact of interface and surface charges on the conduction band of a
heterostructure. The thin solid line represents the conduction band without any bound
charges. However, the difference in polarization of the two semiconductors induces a
positive bound charge at the heterointerface. This pulls the conduction band down
evenly in the thin barrier layer (thick dashed line, which coincides with the thick solid
line in the narrow band-gap semiconductor). This charge attracts electrons on the side of
the narrow band-gap semiconductor. These electrons screen the electric field and, as a
consequence, bend the conduction band upwards. Finally, the polarization field induces
a negative bound charge at the surface, which pulls up the conduction band, as shown
by the thick solid line. ..................................................................................................... 20
Figure 2.7: The lattice constants (in this oversimplified drawing, the atoms that have to
form bonds are represented by red dots) of an unstrained AlGaN are different from that
of an unstrained GaN (left). Yet, when AlGaN grows on the top of GaN, the lattice
constant a, defining the dimensions of the basal hexagon, has to match the lattice
constant of GaN. This exerts a strong stress on the grown layer in the xy-plane. This
stress causes strain (right) in both of the basal plane (ε1) and the growth (ε3) directions,
described by the equations (2.6) and (2.8), respectively. The dashed line represents an
unstrained piece of AlGaN. In piezoelectric materials, the strain results in polarization
field Ppz. In this figure, the Ga-face growth is assumed. In the case of the N-face growth,
the orientation of the polarization would be opposite. a(0) is the lattice constant of an
unstrained GaN or epitaxially strained AlGaN, a(x) the lattice constant of the unstrained
LIST OF FIGURES xii
AlxGa1-xN. The structure in the middle represents the orientation of the wurtzite
structure in the heterostructure. ....................................................................................... 22
Figure 2.8: Piezoelectric polarization of III-N ternaries epitaxially grown on GaN. Dashed
lines with crosses represent linear piezoelectric response to strain, given by the equation
(2.9). The spacing between the crosses represents 10% increase in a binary alloy
fraction. The solid lines represent non-linear response to strain in binary compounds and
then linearly interpolated for the ternaries. The inset shows the impact of non-linearity
of piezoelectric response to strain in the case of AlGaN grown on GaN. For Al fraction
up to 65%, the error is lower than 3%. ............................................................................ 25
Figure 2.9: The closed surface Σ around an interface with abrupt change of polarization,
e.g. AlGaN/GaN interface, is used to calculate the bound sheet charge density by the
equation (2.15). The polarization in the figure is drawn in the negative direction, as it is
true for all tensile and moderate compressive strains (with no or low content of In) in
III-Ns. So, if the magnitude of Pu is larger than that of PL, e.g., in the Ga-face grown
AlGaN on GaN, the bound charge will be positive. ........................................................ 26
Figure 2.10: Vertical cross-section of the conduction band in the heterostructure. The
dotted line is the position of the Fermi level in the semiconductor. Φb is a barrier height.
In the case of a heterostructure or a HEMT far from the gate the barrier height is
determined by the surface sheet charge. Under the gate of a HEMT, it is determined by
the Schottky barrier, modified by the applied gate voltage. Δ is the penetration of the
conduction band edge below the Fermi level at the AlGaN/GaN interface, ΔEC is the
conduction band offset, E0, is the lowest subband level of the 2DEG. The labels
correspond to the ones used in equations (2.17) and (2.18). Adapted from [62]. ........... 28
Figure 2.12: Schematic structure of a HEMT device. The figure is not to scale. In a real
transistor, the length of the device is much larger than the thickness of the AlGaN
barrier. The 2-dimensional electron gas (2DEG) is in the potential well in the GaN
layer, which is the lower band-gap semiconductor in this heterostructure, near the
heterostructure interface. The 2DEG creates the channel, which leads current in the
device. The current flows between the ohmic contacts, the source and the drain, and is
modulated by the voltage applied at the gate, which is a Schottky contact. ................... 30
Figure 2.13: Closing the channel with increasing the negative gate voltage VG. Positive
(negative) voltage applied on the gate lowers (increases) the Schottky barrier Φb. The
solid line represents the conduction band and the dotted line represents the Fermi level.
EF,m is a) A fully open channel where the applied VG is positive, but this is not
necessary. b) Conduction band of a HEMT with negative applied VG, but the channel is
still open. c) The negative VG is larger than the threshold voltage VT and the channel is
closed. .............................................................................................................................. 31
Figure 2.14: The conduction band across the channel. Gradual closing of the channel due
to reverse biasing the gate at a constant drain voltage is shown in the subfigure a). The
x-component of the electric field close to the source is shown in the inset. As shown in
the subfigure b), for small drain voltages, the voltage drop is spread across the whole
channel. However, as the drain edge of the gate becomes depleted, any additional
voltage applied drops in that region and the electric field near the source does not
increase anymore. This is the saturation region. Again, the inset shows the x-component
of the electric field near the source. The drain current is determined by this electric
field. The position of the contacts is indicated at the bottom of each subfigure. ............ 32
Figure 2.15: The effect of surface traps on 2DEG creation, reduction of the negative
surface bound charge and current collapse. a) The polarization charge induces large
electric field in the AlGaN layer. For thin AlGaN barrier, the trap energy level is below
the Fermi level and the states are filled. As the barrier reaches a critical thickness, the
trap level hits the Fermi level and the traps start to empty and become positively
LIST OF FIGURES xiv
charged and, due to the strong electric field, the electrons transfer to the channel
[66,56]. The diagram is only schematic, in reality, the bands in GaN change with the
barrier thickness. For a thin barrier, all electrons are in the traps and not in the channel;
hence the conduction band will be above the Fermi level and will not bend. b) Large
negative VG bias will induce high electric field close to the gate and the electrons form
the gate leak to the empty surface states, create a “virtual gate” [68] and deplete the
channel. In pulsed operation, after the abrupt change of VG, it takes some time to
remove the trapped electrons and for that time the channel remains partially depleted.
This reduces the drain current and hence the expected output power. This phenomenon
is called RF current collapse. .......................................................................................... 34
Figure 3.2: Simulation flow of the Sentaurus simulations platform. Only the tools that were
used in this work are mentioned. Not all input/output files are shown. Each simulation
tool here corresponds to an input file. First, using SSE, files defining the MESH and
DOPING are created. Using these files and the simulation tool SD, the device simulation
(set of voltage sweeps at one of the electrodes) is performed and two types of output
files are produced. CURRENT, that records the current and voltage at all electrodes, and
DATA, that records all pre-specified simulated parameters in all mesh points, e.g. the
electric field, the conduction band, the current density, the electron density, etc. Inspect
and Tecplot SV visualise the CURRENT and DATA files, respectively. ............................ 42
Figure 3.3: HEMT structure created in SSE, showing the mesh. GaN layer is in grey
colour, insulator between the contacts is in red, AlGaN layer between these layers is not
visible in this scale due to its thinness and the density of the mesh in that region. The
three dents on top of the device, represented by the green lines, not originally shown in
the SSE, are the electrical contacts; from left to right it is the source, the gate and the
drain. ................................................................................................................................ 43
Figure 3.4: Detail of the region under the gate of the HEMT device structure shown in
Figure 3.3......................................................................................................................... 43
Figure 3.5: The electron mobility in a GaN HEMT as a function of electron concentration
at three different values of the temperature [105]. The symbols are at VG = 0V. ........... 52
LIST OF FIGURES xv
Figure 3.6: Mobility and velocity as a dependence of the electric field parallel with the
carrier current in the Caughey-Thomas mobility model [106]. The mobility in low
electric field is governed by μn0 and the velocity in high electric field by vsat, hence the
names of the parameters. ................................................................................................. 53
Figure 3.7: Impact of various fitting parameters on ID-VG with focus on the transition from
linear regime to saturation. The parameter that is changed between two simulations is
enclosed in a box. The units of σs are % of what is expected from (2.16) not taking the
surface states into account, as discussed in 2.4.2.1. a) The low field mobility (green line)
and the saturation velocity (blue line) are changed with respect to the simulation
represented by the red line. b) The change of the work-function translates to a shift
along the VG axis. The impact of the surface charge and the contact resistance is
illustrated as well. ............................................................................................................ 55
Figure 3.8: Calibration procedure employing the view of the fitting parameters as operators
shifting the transition region. The ID-VG simulations (dashed lines) are taken from
Figure 3.7 a). The linear and saturation regions are identified (solid thick lines) and
straight lines (solid thin lines) that approximate the selected I-V regions are found. A
crossing-point (empty circles) is calculated for each pair of the straight lines. These
points are then used to calculate by how much we must change the parameters in the
next step to shift the transition point to that of the experimental curve. ......................... 57
Figure 3.9: Schematic illustration of the two HEMT devices simulated in this work (not to
scale). The structure of the devices is 23 nm thick Al0.28Ga0.72N barrier on 1.9μm thick
GaN, which is on top of a SiC substrate (not shown here). The source-drain separation
is 4μm, the source-gate separation is 1μm, and the Ni/Au gate length is 0.25μm. The
electrodes, i.e. the source, the gate and the drain, are represented by thick black lines.
The light orange shading in the GaN layer represents Fe doping. The abrupt change in
the total polarization, spontaneous (2.2.1) and piezoelectric (2.3.2), represented by the
red arrows, gives rise to the bound charge (2.3.3), represented by violet circles with +
and – signs, which induces the 2DEG (2.3.4), represented by the green dashed line,
which makes the operation of a HEMT possible (2.4.1.1). ............................................. 58
the linear region of ID-VG characteristics yields low-field mobility μn0. However, to fit
the ON-current too, it is necessary to change the saturation velocity, which disrupts the
already fitted part of the curve. The same is true for the surface charge. Therefore, it is
necessary to calibrate the device in an iterative process. I-V characteristics for wide
range of gate and drain voltages of the calibrated Device A is shown in Figure 3.12 with
the values reported in Table 3.2. ..................................................................................... 59
Figure 3.11: Calibration process of Device B, follows the procedure described in section
3.4.2. The units of the parameters are cm2.V-1.s-1 for mobility, % of the surface charge
according to equation (2.15), i.e. without the effect of the surface traps, equal to -
3.4×1013cm-2 and ×106cm.s-1 for the saturation velocity. a) initial simulation orig with
estimated fitting parameters and simulations with one parameter changed each, mob
(mobility), surf (surface charge σs) and v_sat (saturation velocity). b) the corresponding
crossing points. c) and d) show the ID-VG characteristics at VD = 3V and VD = 10V,
respectively, calculated from the points in b). As is obvious, the curves differ negligibly
at VD = 3V, which validates the procedure. Using this method, we can generate a
subspace of parameters that give good agreement at a particular voltage and then select
a specific combination based on curves at a different voltage. e) and f) show ID-VG
characteristics after several iterations, simul 31 was selected as the best fit. I-V
characteristics using these parameters for a wide range of VD and VG are reported in
Figure 3.13....................................................................................................................... 60
Figure 3.12: Calibration of Device A. VT ≈ -5V. The parameters used in the calibration are
as follows. The low-field mobility μn0 = 1350cm2.V-1.s-1, the saturation velocity
vsat = 1.04×107cm.s-1, the contact resistance Rc = 350Ω.μm, the Schottky barrier height
ΦSch = 1.2eV, the charge at the interface σi = 1.28×1013cm-2 and the charge at the
surface σs = -4.76×1012cm-2. ............................................................................................ 61
Figure 3.13: Calibration of Device B. VT ≈ -5.2V. The parameters used in the calibration
are as follows. The low-field mobility μn0 = 1450cm2.V-1.s-1, the saturation velocity
vsat = 9.53×106cm.s-1, the contact resistance Rc = 750Ω.μm, the Schottky barrier height
ΦSch = 0.95eV, the charge at the interface σi = 1.28×1013cm-2 and the charge at the
surface σs = -3.65×1012cm-2 ............................................................................................. 62
Figure 3.14: ID-V characteristics of Device A. a) The measured dependence of the drain
current ID on the drain and gate and voltage, VD and VG. (Figure 3.12 (a) and (b)
LIST OF FIGURES xvii
Figure 3.15: ID-V characteristics of Device B. a) The measured dependence of the drain
current ID on the drain and gate voltage, VD and VG. (Figure 3.13 (a) and (b) combined).
Calibration error of the simulated ID, expressed in % above/below the measured values,
(b) as a function of both VD and VG and separately, as a function of (c) VG and (d) VD.. 66
Figure 3.16: The calibration error as a dependence of a) VG and b) VD, averaged over a
range of simulated VD (1 – 10 V) and VG (-4 – 1 V) points, respectively. The error for
VG = -5V (almost 25% for Device A), since it is close to VT, is excluded from the
calculation in subfigure b). The average error for all values of VD is below 6%. For VG
between -3 and 0V in Device A and between -1 and 2V in Device B, the average error is
below 2.5% and 3.5%, respectively. ............................................................................... 67
Figure 4.1: Schematic representation of the converse piezoelectric effect, based on the set
of equations (4.1). The crystal direction is shown on the left. The wurtzite structure is
asymmetrical in the z direction, so the z-component of the electric field, Ez, will have
opposite effect in the positive and negative directions. The two rows represent the
opposite directions of Ez. Other components of E induce only shear strains. In all
pictures, the thick grey lined square represents the crystal without the effect of E; the
thin black lined square represents the deformation due to the E, as well as external
stresses needed to maintain the desired shape. The three columns represent three
different boundary conditions. The first column is in the absence of any mechanical
external forces, σi = 0, for all i. In this case, the crystal simply expands in one direction
and contracts in the other. The second column is for a completely fixed structure,
without the possibility to change the shape. In this case, E will produce forces on
surrounding material. The arrows are in the direction of external forces that need to be
applied on the structure to prevent it from deforming. Finally, the clamped model, given
by the equations (4.3) and (4.4), is shown in the third column. The dashed line
represents a freestanding structure, the grey line represents a strained structure, e.g. a
thin AlGaN barrier layer grown on GaN, strained to match the underlying layer, and the
black line represents the structure under the impact of E. The stress shown in this
column is a change in the stress already present due to the layer being already strained.72
LIST OF FIGURES xviii
Figure 4.2: Subfigures a) and b) show vertical strain ε3 (4.3) and lateral stress σ1 (4.4)
respectively, in accordance with the clamped model, as a function of the z-component of
the electric field E3, for various Al fractions x of AlGaN, assumed to grow on a thick
relaxed GaN layer. From the subfigure a) it may seem that an electric field parallel with
the z direction (E3 > 0) reduces the strain. Nevertheless, we must remember that, the
layer grows with a built-in lateral strain ε1; the vertical strain ε3 is a result of the
assumption of no force applied in the z direction. Since the electric field in any direction
exerts an additional force, any variation in strain will only increase the total strain of the
crystal. This is demonstrated by the subfigure c) in which strain energy per unit volume
versus the vertical strain or electric field is plotted. A non-zero electric field can
increase or decrease the vertical strain, but it always increases the strained state of the
crystal. ............................................................................................................................. 74
Figure 4.3: The deficiency of the clamped model comes from its simplifying assumptions
is shown by considering two adjacent elements of a piezoelectric material. Dashed lines
represent the situation before applying the electric field. The electric field in the
“element 2” is greater than that in the “element 1”. From (4.3) and (4.4) it follows that,
the lateral stress and vertical strain in those two elements will be different. Two obvious
problems arise from this result. One is non-equilibrium in stress; the “element 2” will
press on the “element 1” with larger force than the other way around. The second is that
a point on the top boundary of the two elements (full circle) will split in two under the
influence of the electric field. In a solid matter, this is not possible. As a consequence,
even by neglecting (or in the absence of) the x and y components of the electric field,
there will still be shear strains and stresses in the device. ............................................... 75
Figure 4.4: The impact of the converse piezoelectric effect on the bound sheet charge at the
heterojunction interface. The situation on the left is without taking the effect into
account. The spontaneous polarization is not shown, because it is not affected by the
electric field or the additional strain. The gray square represents the unstrained AlGaN.
AlGaN grows with strain on the relaxed GaN, which results in piezoelectric polarization
in AlGaN, and according to (2.15), the difference in the polarizations leads to formation
of the interface charge. The charge induces the electric field E and, via the converse
piezoelectric effect, produces additional strain in both AlGaN and GaN. This modifies
the polarization in both layers and hence alters the bound charge, as derived in (4.9).
LIST OF FIGURES xix
The bias applied at the electrodes, especially the gate, modifies the electric field and
therefore the bound interface charge. .............................................................................. 76
Figure 4.5: Simulation flow to show the impact of the converse piezoelectric effect on the
simulated I-V characteristics and electric field distribution in the device. ..................... 78
Figure 4.6: Distribution of the electric field in the device at drain voltage V DS = 3V and
gate voltages VGS = -6V and VGS = 0V, just under the threshold. .................................. 79
Figure 4.7: The electric field E distribution near the gate in the GaN layer, 0.1 nm under
the interface, in the linear regime, VD = 3V, and saturation, VD = 20V. The position of
the gate is indicated by the thick solid line at the bottom of each graph and vertical thin
dashed lines on the edges of the gate. ............................................................................. 79
Figure 4.8: The distribution of change of the vertical strain in the region close to the gate in
the off-state and at a low drain voltage, VD = 3V, VG = -6V. ......................................... 81
Figure 4.9: The distribution of change of the vertical strain in the region close to the gate in
the off-state and at a high drain voltage, VD = 20V, VG = -6V. ...................................... 81
Figure 4.10: The distribution of change of the vertical strain in the region close to the gate
in the on-state and at a high drain voltage, VD = 20V, VG = 0V. .................................... 81
Figure 4.11: Distribution of the polarization induced charge in the device, close to the gate,
at VDS = 3V. ..................................................................................................................... 82
Figure 4.12: Distribution of the polarization induced charge in the device, close to the gate,
at VDS = 20V. ................................................................................................................... 82
Figure 4.13: Simulation of ID-VG characteristics (top) in the linear region of the device, at
VD = 3V. The uncoupled simulation is without the contribution of the converse
piezoelectric effect. The simulation labelled as coupled (space) takes the contribution of
the bound space charge into account, and the simulation labelled as coupled takes also
the modification of the bound sheet charge at the interface into account. The two bottom
graphs show the drain current shift in the coupled (space) (left) and coupled (right)
models, with respect to the gate voltage VG. In saturation regime, the contribution of
both effects is comparable, while for the linear regime and at the threshold voltage, the
modification of the bound sheet charge at the interface is dominant. ............................. 83
LIST OF FIGURES xx
Figure 4.14: Modification of the bound sheet charge σb at the interface. Away from the
gate, the electro-mechanical coupling results in a constant reduction (approx. 1.5%) of
the charge, irrespective of the voltage applied. Under and close to the gate, the electric
field is strongly modified by the voltage applied at the electrodes, especially at the gate.
Therefore, the bound sheet charge is modified as well. .................................................. 85
Figure 4.16: The electric field E distribution along the channel, under or close to the gate,
0.1 nm under the heterojunction interface. The top subfigures show the electric field
change at low drain voltage, the bottom subfigures at high VG. Ex is shown on the
subfigures on the left, Ez on the right. The effect of the polarization induced bound
space charge alone (top: coupled (space)) on the simulation of E is negligible. At gate
voltages below threshold (left), the change of Ex is independent of VD. ......................... 86
Figure 4.17: The conduction band (CB) in the channel, 0.1 nm under the heterojunction
interface. The top subfigures show the CB below the threshold voltage of the device,
where the potential barrier prevents electrons from flowing through the channel and
hence the channel is closed. For both of the considered drain voltages, the barrier is
increased. In the bottom subfigure, the CB shift due to the converse piezoelectric effect
for various bias conditions is summarized. For voltages below and around the threshold
voltage, the region that has an impact on the device operation is under the gate, since it
determines the height of the potential barrier. For voltages above the threshold, the
important region is between the source and gate. With increasing the gate voltage, the
CB shift in this region fades away................................................................................... 87
Figure 5.1: c) An illustration of the dependence of the measured current on the quiescent
bias at which the device is hold before the actual measurement. The real a) ID-VG and b)
ID-VD characteristics for two different quiescent bias voltages. The term open-channel
refers to the quiescent bias of VG,q = 0V, VD,q = 0V; whereas the measurement that
showed the current collapse to VG,q = -4V, VD,q = 25V. The dots represent the
measurement point of VG = 0V, VD = 5V, but the measured current depends on the
LIST OF FIGURES xxi
quiescent bias, a bias applied before the measurement. This phenomenon is referred to
as DC-RF dispersion or “knee-walkout” due to its representation in the I-V plane (blue
arrow), as mentioned in section 2.5.2. The I-V characteristics for a degraded device in
the open-channel condition is shown in blue lines. ......................................................... 91
Figure 5.2: The mechanism responsible for the current collapse. Due to the strong electric
field (red arrow), through the means of the Poole-Frenkel conduction (green arrow), the
electrons leak to the surface of the device and electrostatically deplete the channel (blue
arrow) and hence cause reduction of the saturation current. This is the primary Poole-
Frenkel mechanism. The secondary mechanism is that the electrons at the surface create
additional electric field which forces the later leaked electrons to transfer to the traps in
the bulk and to the AlGaN/GaN interface. During a stress test, in the region of the high
electric field (red arrow), new dislocations, and therefore traps, are created. This leads
to a permanently degraded device and reduced current in subsequent measurements.
Adapted from [125] ......................................................................................................... 92
Figure 5.3: A scheme of the device with focus on the regions (red lines) with the trapped
charge to reproduce the pulsed measurements of the I-V characteristics from the class
AB point. ......................................................................................................................... 93
Figure 5.5: (a) ID-VG and (b) ID-VD characteristics for different values of surface electron
density, 1-3×1013cm-2, placed at the edges of the gate, in a region extending 50nm
towards the other electric contacts. Increasing the amount of charge leads to reduction
of the saturation current, but the slope of the linear regime of ID-VD remains unchanged.
(c) The impact of the surface charge on the transconductance of the device, the black
line represents the device with no trapped charge. .......................................................... 94
Figure 5.6: The impact of charge in regions of two different sizes (to 50nm and 150nm
away from the gate) on the I-V characteristics (a, b) is compared. Several values of the
LIST OF FIGURES xxii
electron sheet density are used for both regions. (c) Transconductance for the 150nm
region of the trapped charge (for 50nm region it is reported in Figure 5.5 c). ................ 95
Figure 5.7: (a) ID-VD and (b) ID-VG characteristics for different values of the extent of the
electron charge from the gate, 30-70nm, of constant sheet density, 2×1013cm-2. The
impact of extending the region of the trapped charge and increasing the charge density,
shown in Figure 5.5, is similar, but not equal, as (c) the transconductance dependence
shows. .............................................................................................................................. 96
Figure 5.8: The sole effect of charge trapped at the AlGaN/GaN interface under the gate is
the threshold voltage, VT, shift. The slope and the ON-current remain unchanged. Apart
from the varying charge under the gate, there was sheet charge density of 6.5×1012cm-2
extending to 150 nm on both sides of the gate, reported in Figure 5.6. .......................... 96
Figure 5.9: Asymmetric charge distribution around the gate. The impact of the trapped
charge at the source and drain sides of the gate is investigated separately, for low charge
density in a large region. (a) ID-VG and (b) ID-VD characteristics and (c)
transconductance are compared....................................................................................... 97
Figure 5.10: Asymmetric charge distribution around the gate. The impact of the trapped
charge at the source and drain sides of the gate is investigated separately, for high
charge density in a small region. (a) ID-VG and (b) ID-VD characteristics and (c)
transconductance are compared....................................................................................... 98
Figure 5.11: The best fit of pulsed I-V characteristics using uniform blocks of trapped
charge placed asymmetrically around the gate. While achieving reasonable fit of ID-VD,
as well as VT, and ON-current, the transition between the linear regime and the ON-
current in ID-VG (blue ellipse) remained problematic. This appeared for all simulated
values of VD, from 1V to 10V (not reported here). ......................................................... 99
Figure 5.12: a) Schematic illustration of the exponential charge distribution model, which,
via the parameters A, B and , allows for independent control of the sheet charge
density at the gate edge QS(0), on the front of the distribution QS(d) and the total charge
QL trapped at the surface on either side of the gate. The method of calculation of the
parameters A, B and is described in the next subsection, 5.4.2, and given by equations
(5.3) and (5.4). b) Schematic illustration of the arrangement of the charge distribution in
LIST OF FIGURES xxiii
the device. Exponential distribution on the source and drain sides of the gate and a
constant sheet charge density under the gate. ................................................................ 100
Figure 5.13: a) ID-VG and b) ID-VD calibration of current collapse measurement, from class
AB point, at VD = 3V and at VG = 0V, respectively. Subfigure c) shows the impact of
the surface trapped charge on transconductance, where the simulations without and with
the current collapse are represented by thin and thick lines, respectively. The colours are
consistent with a). Figure 5.14 shows the corresponding surface electron distributions at
the source and drain sides of the gate. ........................................................................... 104
Figure 5.14: The surface charge distribution obtained by calibrating the pulsed I-V
characteristics, shown in Figure 5.13. The parameters of these distributions are
summarized in Table 5.2. .............................................................................................. 105
Figure 5.15: a) The simulated ID-V characteristics during the current collapse (Figure 5.13
a) and b) combined). The error of the calibration with respect to the measured
characteristics (squares in Figure 5.13) (b) as a function of both VD and VG and
separately, as a function of (c) VG and (d) VD. .............................................................. 106
Figure 5.16: The lateral stress σx in the device at VG = -4V and VD = 25V as calculated
from (4.4) within the clamped model. The bottom subfigure offers a few cross-sections
of the parameter. The value y in the top subfigure indicates the distance from the
interface while in the bottom subfigure it indicates the distance from the surface. The
gate is indicated as a pink line in the top subfigure and by the vertical lines in the
bottom subfigure............................................................................................................ 108
Figure 5.17: I-V simulations using the constant charge density model. The impact of the
varying charge density ρ0 was investigated, while the threshold stress σ0 was kept
constant. ......................................................................................................................... 109
Figure 5.18: I-V characteristics for various values of the threshold stress and trapped
charge density. The third value in the legend is the total charge introduced in the device
in each particular simulation. ........................................................................................ 109
Figure 5.19: Simulated I-V characteristics using the linear dependence model for various
values of the parameters. Note that the total charge introduced in the device in the blue
simulation is lower than in the green simulation, yet the impact on the reduction of the
LIST OF FIGURES xxiv
drain current is stronger. Due to higher threshold value, the region where the charge is
introduced is smaller. .................................................................................................... 110
Figure 6.1: At high electric field E at the edge of the gate, electrons leak to the surface (red
line). Then, due to the strong E, they are transported away from the gate (blue arrows).114
Figure 6.2: The magnitude of the x-component of the electric field at the surface of a
HEMT device with respect to the distance from the gate edge. .................................... 116
Figure 6.3: The dependence of the surface current density, j, on the x-component of the
electric field for various values of parameters m and b, calculated using equations (6.3)
and (6.4). The values of E at the drain (ED = 3.51 MV/cm) and source edge
(ES = 0.84 MV/cm) of the gate, at VG = -4V and VD = 25V in the simulated device, were
taken for E1 and E2. The value of jD was set to 45 (solid lines) and
30 (dashed lines), the ratio jD/jS was set to 5 (blue), 10 (green),
and 15 (red). .................................................................................................................. 116
Figure 6.4: The dependence of the emission frequency, with which an electron tunnels
from a surface trap to the next trap, on the electric field, according to the Poole-Frenkel
transport model, expressed by equation (6.5). ............................................................... 118
Figure 6.5: The velocity of an electron travelling across a HEMT surface travelling from
trap to trap, as given by equation (6.6). ......................................................................... 118
Figure 6.6: Profiles of the electron density on the device surface at the drain side of the
gate. The gate edge is positioned at x = 1.25m. All four graphs show four different
values of initial surface currents on the drain edge of the gate, j D,0, namely 50, 45, 40
and 35 e-m-1s-1, represented by thick solid, dotted, thin solid and dashed lines,
respectively. The ratio between initial surface currents at the drain and source sides is in
all cases jS,0/jD,0 = 5. The top row shows the distribution at two specific moments, using
(a) slow and (b) moderate velocity profiles. The bottom row compares the distributions
that result from different velocity profiles, at a specific time of the simulations. c) Slow
and moderate at 400s, and (d) slow, moderate and fast at 160s. .............................. 120
Figure 6.7: a) The electric field distribution using slow and moderate velocity profiles after
160s and 80s, respectively. The black line is the electric field before the electron
leakage. Higher leakage current and hence higher electron density (solid lines) leads to
LIST OF FIGURES xxv
higher electric field and therefore higher velocity at the front of the electron
distribution. This results in the electrons getting farther from the gate. The colouring
corresponds to the Figure 6.6 c). b) The electric field (arrows) before (top) the leakage
and the electric field (blue and green) associated with the leaked electrons (bumps) and
its effect on the total electric field (red arrows with blue and green outlining). Higher
electron concentration increases the electric field and velocity at the front of the stream
of electrons and reduces it at the end close to the gate. The vertical black line represents
the position of the gate edge. Note: The bump on the graph of Ex in the figure on the left
is due to an abrupt change in mesh spacing in the simulator. ....................................... 121
Figure 6.8: The electron distribution on (a) the source and (b) drain sides of the gate after
160s (solid) and 400s (dashed). The simulations were done for all three above
mentioned velocity profiles, slow (blue), moderate (red) and fast (green). b) (drain side)
also shows the impact of changing the jD,0/jS,0 ratio (light coloured lines). The effect is
stronger for slower velocities. Using the fast velocity profile, the electrons on the drain
side crossed almost 1m in just 160s. The gate is positioned between x=1m and
x=1.25m. ..................................................................................................................... 122
Figure 6.9: The electric field at the drain (a) and source (b) edges of the gate. Red lines
show simulations with same electron leakage parameters (jD,0, jD,0/jS,0) and different
velocity profiles, given by different combinations of ΔG and s. Solid lines show
simulations with slow electron velocity, the same jD,0/jS,0 ratio, but varying initial
surface current density jD. On the drain side, pink and orange dashed lines represent
simulations with the same jD as the red line, but varying jD,0/jS,0 ratio. ......................... 123
Figure 6.10: Due to the electric field at the edge of the gate, the electrons tunnel from the
gate to the device surface. a) Electron surface current density on the drain edge jD given
by the Frenkel-Poole emission model. b) The total charge leaked to the surface. The
colour coding is the same as in Figure 6.9, except for the fast velocity profile in b),
where the dotted red line is replaced with solid black, to make the graph more readable.123
Figure 6.11: Ex at the drain edge of the gate simulated for various values of jD,0, using two
different velocity profiles and two different electron leakage profiles, with one
combination missing, for the sake of clarity of the figure. This is a follow-up to
Figure 6.9 left, for simulations with higher jD,0, with the highest jD,0 from that figure
LIST OF FIGURES xxvi
reprinted in this one with grey lines. Follow-up to this figure with even higher jD,0 is
Figure 6.13. Increasing jD,0 further reduces the electric field even faster. For
jD,0 = 2,000 e-m-1s-1, the electric field at the gate edge drops to half its initial value
in less than 6 s. This has a huge impact on the development of the electron
distribution. The corresponding electron distributions are presented in Figure 6.17: solid
lines (ΔG = 0.3eV, jD,0/jS,0 = 5) – top left; dashed lines (ΔG = 0.3eV, jD,0/jS,0 = 20) – top
right; dotted lines (ΔG = 0.25eV, jD,0/jS,0 = 5) – bottom left; unreported here
(ΔG = 0.25eV, jD,0/jS,0 = 20) – bottom right. Figure 6.20 right compares the distributions
(top) and jD (bottom) simulated with the same value of jD,0 = 2,000 e-m-1s-1 (red
lines here). The corresponding jD is shown in Figure 6.12. .......................................... 127
Figure 6.12: The dependence of the temporal evolution of jD on jD,0, on the emission model
parameters and on the transport velocity of the electrons. Higher jD,0 results in faster
reduction of jD both relative to jD,0 (left) and in absolute values (right). QL, which is the
integral of jD, for ΔG = 0.3eV is reported in Figure 6.22 and Figure 6.23 for jD,0/jS,0 = 5
(solid lines) and jD,0/jS,0 = 20 (dashed lines), respectively. The corresponding Ex is
shown in Figure 6.11, all other parameters as described therein. ................................. 127
Figure 6.13: Ex at the drain edge of the gate, for jD,0 = 5,000 e-m-1s-1 and more, which is
a follow-up figure to Figure 6.11 (with one of the simulations shown in that figure
reprinted here). As expected, further increase in jD,0 speeds up the reduction of Ex at the
gate edge and hence the electron tunnelling to the device surface. For the parameters
jD,0 = 20,000 e-m-1s-1 and jD,0/jS,0 = 5, Ex drops to half its initial value in 0.47s, and
for jD,0 = 50,000 e-m-1s-1 and jD,0/jS,0 = 20 in 0.25s. Some of the corresponding
electron distributions are reported in Figure 6.18 (jD,0/jS,0 = 5), Figure 6.19 (jD,0/jS,0 = 20)
and Figure 6.21 (jD,0 = 20,000 e-m-1s-1). The corresponding jD is shown in
Figure 6.14..................................................................................................................... 128
Figure 6.14: Further increase of jD,0 causes further acceleration of the reduction of jD (left)
which, for high jD,0 quickly drops below the jD of the lower jD,0 (right). The higher is
difference between jD,0 of two simulations, the faster this happens (right). QL is shown
in Figure 6.22 and Figure 6.23 for jD,0/jS,0 = 5 (solid lines) and jD,0/jS,0 = 20 (dashed
lines), respectively. The corresponding Ex is shown in Figure 6.13, all other parameters
as described therein. Note: the slight oscillation in the simulation of jD for
LIST OF FIGURES xxvii
jD,0 = 20,000 e-m-1s-1 and jD,0/jS,0 = 5 (solid black line) is caused by the fact that the
time step was insufficiently short. This will be expanded on in the next section (6.5.3).128
Figure 6.15: Ex at the source edge of the gate. The corresponding distribution and jS for
jS,0 = 100 e-m-1s-1 (orange lines), compared with jD,0/jS,0 = 20 (not shown in this
figure) is shown in Figure 6.20, top and bottom left, respectively. The jS corresponding
to this figure are shown in Figure 6.16. ......................................................................... 130
Figure 6.16: jS for various jS,0 and various velocity profiles. QL for ΔG = 0.3eV is reported
in Figure 6.24 left; the corresponding Ex in Figure 6.15; and other parameters as
described therein. The cause of the oscillations for jS,0 = 1,000 e-m-1s-1 (solid light
green line) is a too large time step. ................................................................................ 130
Figure 6.17: The electron distributions at the drain side of the gate for four different values
of jD,0 at two time points, 20s (dotted lines) and 100s (solid lines). Figures shows
simulation results for slow (top) and moderate (bottom) velocity profiles, and two
electron emission profiles, jD,0/jS,0 = 5 (left) and jD,0/jS,0 = 20 (right). The distributions
from simulations for jD,0 = 2,000 e-m-1s-1 are compared in Figure 6.20, right. Ex and
jD for all simulations, except the one on bottom right, are reported in Figure 6.11 and
Figure 6.12; QL for top left (ΔG = 0.3eV, jD,0/jS,0 = 5) and top right (ΔG = 0.3eV,
jD,0/jS,0 = 5) are shown in Figure 6.22 and Figure 6.23, respectively. ........................... 131
Figure 6.18: A follow-up to Figure 6.17 top left, the electron distributions for three
different values of jD,0 at three time points, 2s, 10s and 30s. In spite of grand
differences in distributions shortly after the start of the leakage, at 2s (dotted lines),
due to unequal drop in Ex and hence jD, the distributions converge to roughly the same
“shape” later, at 30s (solid lines). One of two main differences among the distributions
is the maximal distance the electrons reached, which changes less and less with higher
values of jD,0 (compare with Figure 6.17). The other is that with higher jD,0, the electron
density at the gate edge is higher (right). The corresponding Ex and jD are reported as
solid lines in Figure 6.13 and Figure 6.14, respectively; the corresponding QL is shown
in Figure 6.22. ............................................................................................................... 132
Figure 6.19: A follow-up to Figure 6.17 top right, the electron distributions for very high
jD. As in true for simulations in Figure 6.18 (jD,0/jS,0 = 20), higher jD means more
electrons at the gate edge and the front of the distribution is further away, although the
LIST OF FIGURES xxviii
latter is less significant for very high jD. Otherwise, the resulting distributions are
similar. The corresponding Ex and jD are reported as dashed lines in Figure 6.13 and
Figure 6.14, respectively; the corresponding QL is shown in Figure 6.23. ................... 133
Figure 6.20: The electron distributions (top) and jS as a function of time (bottom) for the
source (left) and drain (right) sides of the gate. This figure compares simulations with
the same jD,0 (jS,0), for two different velocity profiles (given by the parameter ΔG) and
two different electron emission parameters (jD,0/jS,0). Slower velocity and lower jD,0/jS,0
both mean more electrons close to the gate and hence a distribution closer to the “ideal”.
On the drain side, the corresponding Ex and jD are reported as red lines in Figure 6.11,
Figure 6.12, respectively. Blue and green lines in this figure are results of simulations
with the same parameters for both the source and the drain sides. The distributions (top)
are a higher jD follow-up to Figure 6.8. ......................................................................... 133
Figure 6.21: While Figure 6.18 and Figure 6.19 compare the electron distributions for
varying jD,0, while keeping jD,0/jS,0 constant, this figure compares the impact of varying
jD,0/jS,0 at jD,0 = 20,000 e-m-1s-1. It is a higher jD follow-up to Figure 6.20 top right
for ΔG = 0.3eV. As, keeping the previous results in mind, one would expect, the
difference between these simulations reduces with time and the main difference is the
electron density close to the gate. The corresponding Ex and jD are shown in Figure 6.13
and Figure 6.14, respectively. ....................................................................................... 134
Figure 6.22: The total amount of the electrons leaked to the drain side of the gate for
various jD,0 at jD,0/jS,0 = 5, for slow velocity profile. The higher is the initial leakage, the
more electrons leak to the surface. However, since jD falls rapidly for simulations with
high jD,0 (Figure 6.12 right and Figure 6.13 right), the increase of QL , after the initial
burst, reduces faster for higher jD,0. This is well demonstrated in the figure on the top
right. The corresponding electron distributions are reported in Figure 6.17 top right and
Figure 6.18..................................................................................................................... 134
Figure 6.23: QL for various jD,0 at jD,0/jS,0 = 20, for slow velocity profile. The corresponding
electron distributions are reported in Figure 6.17 top left and Figure 6.19. .................. 135
Figure 6.24: The electrons leaked to the surface side of the gate for jD,0/jS,0 = 5 (left) and
jD,0/jS,0 = 20 (right). ........................................................................................................ 135
LIST OF FIGURES xxix
Figure 6.25: The impact of the time step on the simulations, here exemplified using the
physical parameters jS (left) and jD (right). The parameters used for these simulations are
as follows. Top: jD,0 = 50,000 e-m-1s-1, jD,0/jS,0 = 5; Centre: jD,0 = 50,000 e-m-1s-1,
jD,0/jS,0 = 20; Bottom: jD,0 = 75,000 e-m-1s-1, jD,0/jS,0 = 20. The difference in the time
the electric field was kept constant (frozen) tf was also different in these simulations.
The decrease in tf was two-fold (top), five-fold (centre) and twenty-fold (bottom). The
higher is the difference in tf, the more the simulation is changed. ................................ 137
LIST OF TABLES xxx
List of Tables
Table 1.1: Johnson‟s figure of merit (JM) [3] and related material parameters for different
materials. Values taken from [4,2], except for a) reference [5] b) reference [6] c)
reference [7] d) reference [8]............................................................................................. 2
Table 2.2: Voigt notation for reducing the number of indices of a property that is
symmetrical in two of its indices. .................................................................................... 11
Table 2.4: Material parameters that determine heat dissipation by the substrate. a)
Reference [28], b) Reference [74]. .................................................................................. 35
Table 3.1: Fitting parameters used in the simulations of Device B, reported in Figure 3.11.
The best fit was achieved in simul 31, printed in bold font............................................. 61
Table 3.2: The values of the fitting parameters in the calibrated devices and in the
literature........................................................................................................................... 64
Table 5.1: List of values of the parameters used in the calibration of the exponential charge
model. The top table represents the first set of values, combinations of which were
simulated. Based on the results from these simulations, some values were swapped for
new ones, shown in the bottom table. The new values are printed in bold. QS (cm-2) is
sheet charge and QL (cm-1) is total charge of the distribution. ...................................... 103
LIST OF TABLES xxxi
Table 5.2: Parameters of the exponential distribution that yielded the best agreement with
experimental data, as shown in Figure 5.13. QS (cm-2) is sheet charge and QL (cm-1) is
total charge of the distribution at the specified side of the gate. The surface electron
distributions are visualized in Figure 5.14. ................................................................... 105
Table 6.1: The values of physical parameters, that enter the Poole-Frenkel transport model,
used in the simulations. ................................................................................................. 118
Table 6.2: List of simulations analysed in this section. tf is the time step, for which the
electric field was frozen in each simulation. Where there are two numbers, two
simulations were performed, to analyse the impact of the time step on the evolution of
the electric field and the resulting electron distribution. The colours of the lines
represent the colours used in the following figures for the respective simulations. ..... 125
1 Introduction
1.1 Background
The semiconductor industry has been dominated by silicon technology for decades with its
established CMOS process. The major driving force for the growth of the industry was
scaling. 22-nm node technology is currently in development and 11-nm node is predicted
to reach the market in 2015 [1]. In this or the following decade, further scaling will reach
the limits. Hence, new semiconductor materials, SiGe, SiC, III-V and II-VI, and new
transistor architectures are investigated to replace or to complement silicon.
listed for the above mentioned materials in Table 1.1, along with other related material
parameters. JM is a product of the breakdown electric field and saturation velocity. By
comparing the JM values, it becomes clear that the most promising materials for high
power, high frequency applications are SiC and GaN. However, with respect to fabricating
transistors, GaN has the advantage over SiC of forming heterojunctions.
Table 1.1: Johnson‟s figure of merit (JM) [3] and related material parameters for different materials. Values
taken from [4,2], except for a) reference [5] b) reference [6] c) reference [7] d) reference [8].
Apart from the large band-gap and favourable JM, the material properties of N-based
semiconductor compounds furnish also other advantages over the conventional III-Vs,
leading to superior performance parameters. The presence of spontaneous polarization and
the fact that the piezoelectric polarization is approximately ten times higher than in As-
based semiconductors, give rise to polarization doping in N-based heterostructures, first
predicted by Bykhovski et. al. [9], which makes the elemental doping of the N-based
devices unnecessary. From the absence of doping follows the reduction of ionized impurity
scattering and therefore increase in electron mobility. The resulting two-dimensional
electron gas (2DEG) density in GaN-based devices reached above 1013cm-2, making it five
times larger than that of GaAs. The AlGaN/GaN 2DEG was for the first time observed by
Asif Khan et. al. in 1992 [10], followed by the first demonstration of promising DC [11]
and RF [12] performance of AlGaN/GaN HEMTs. Nowadays, the achieved output power
density of 30-40W/mm is more than ten times higher than that of GaAs based transistors.
As a consequence, to achieve the same output power, the size may be ten times reduced,
reducing the cost of the device [13].
1.1 Background 3
With respect to applications in optoelectronic devices, the advantage of GaN over SiC rests
on the direct band-gap, allowing for higher intensity of the emitted light, and the advantage
over GaAs rests on larger band-gap, allowing for LEDs and LDs with shorter wavelength.
The GaN/InGaN/GaN double heterostructure was first used to fabricate efficient blue light
LED in 1993 [14] and LD in 1997 [15] by Nakamura. Since then, GaN was also utilized
for fabricating UV detectors [16].
Figure 1.1: Summary of failure mechanisms in AlGaN/GaN HEMTs. Taken from [17].
collapse, electrons are trapped mainly on surface states. It is proposed that the mechanism
responsible for the electron leakage from the Schottky contacts to the traps is Poole-
Frenkel electric field dependent surface conduction model [32,33,34,35], first described in
1938 [36].
In this thesis, we aim to develop numerical simulations methodology for studying the
failure mechanisms related to trapped charge, i.e. current collapse and device degradation.
This task includes the following objectives:
(ii) To develop a self-consistent methodology for including the impact of the field
induced polarization on the transistor characteristics and the accuracy of the
calibration process.
(iii) To study the current collapse in GaN HEMTs and to develop an automated
procedure for extracting the distribution of the related trapped charge that would
reproduce the effects as measured by I-V characteristics.
(iv) To explore the non-linear transport of carriers injected from the gate to the
HEMT surface and the following trap-to-trap hopping to reproduce the surface
charge distribution obtained by the current collapse calibration.
1.3 Outline
Chapter 2 introduces the crystal, electrical and elastic properties of III-N materials,
discusses the polarization effects and the origin of the resulting 2DEG. Furthermore, N-
based heterostructures and their properties, such as band discontinuity and polarization
induced bound charge, are discussed. Finally, the operation of AlGaN/GaN HEMTs is
explained, with referral to state of the art devices and key challenges in the technology.
Chapter 3 describes the methods and tools used in this work, i.e. the commercial simulator
Sentaurus by Synopsys and scripts developed to manipulate the input files and to
automatically perform and evaluate simulations. Besides that, the model with its basic
equations governing the simulations is described. Finally, the calibration of the numerical
simulations against the experimental data is explained in detail and exemplified in respect
of the two transistors used later on in this study.
Chapter 5 investigates trap related phenomena, namely current collapse, more specifically
DC-RF dispersion or “knee” walkout, and device degradation. Since the surface trapped
charge plays role in both phenomena, the impact of various distributions on I-V
characteristics is investigated, to build on the gained insights later. The current collapse
experimental data is calibrated with an asymmetrical exponential surface charge
distribution. Further, the electric field is linked to mechanical stress generated in the device
and that stress, in turn, to defects with trapped charge. Two relationships between the stress
in the device and the trapped charge are proposed and investigated through their impact on
the I-V characteristics.
Using the surface charge distribution obtained by the current collapse measurement
calibration in Chapter 5 as a target, Eyring‟s reaction rate model and Poole-Frenkel
emission model are employed to reproduce the distribution in Chapter 6. Wide range of
input parameters is considered to investigate the temporal evolution of several quantities,
1.3 Outline 6
such as the electron surface distribution, the electric field at the surface and at the gate
edges and the surface current densities at the gate edges.
Chapter 7 summarizes the results obtained in previous chapters and draws the
corresponding conclusions. It also outlines future challenges in GaN HEMT reliability
simulations.
2 GaN and Related Devices
This chapter provides background information about GaN as a semiconductor material and
its implementation in the design of high electron mobility transistors (HEMTs) for high
frequency – high power applications. Section 2.1 outlines the physical properties of GaN
as a semiconductor material. Section 2.2 deals with its polarization properties that play
important role in the formation of 2DEG at the AlGaN/GaN interface. The properties of
the 2DEG are discussed in more details in Section 2.3. Section 2.4 introduces the basic
concepts and operation of AlGaN/GaN HEMTs while Section 2.5 discusses some of the
challenges of the GaN technology.
III-N (AlN, GaN, InN) semiconductors crystallize in both the wurtzite hexagonal close
packed (HCP) (α-phase) and cubic zinc-blende (β-phase) crystal structures. The wurtzite is
the more stable structure and possesses the spontaneous polarization Psp, which can be
exploited in creating high-density 2DEG (two dimensional electron gas) at III-N
heterointerfaces. This crystal structure is therefore the structure of choice for device
production, and hence will be of interest in this work. On the contrary, the conventional
III-V semiconductors, such as GaAs or InP, crystallize in the zinc-blende structure.
2.1 Physical Properties 8
Figure 2.1: III-Nitrides crystallize in cubic zinc-blende (left) and hexagonal wurtzite (right) structure. They
both lack the centre of symmetry, so they show piezoelectricity. The wurtzite, which is a lower symmetry
crystal, possesses also the spontaneous polarization. The lattice parameters a, c and u are shown for the
wurtzite structure. The arrows on the zinc-blende structure depict the set of <111> directions. Adapted from
[37].
The wurtzite structure lacks inversion symmetry along its c-axis (called the pyroelectric
axis) and hence the directions [ 0001 ] and [ 0001 ] are not equivalent, as shown in
Figure 2.2. This lack of symmetry gives rise to the spontaneous polarization in III-N
semiconductors. The conventional positive direction of the c-axis in III-Ns is the one that
follows the direction from the group III atom to the N atom. According to the atom on the
top position of the {0001} bilayer, the ( 0001 ) plane is called the Ga- (Al-, In-) face, while
the plane ( 0001 ) is called the N-face. The crystal is then said to have a Ga-(Al-, In-)
2.1 Physical Properties 9
polarity or N-polarity. The electric dipole in III-Ns is directed from the N to the Ga (Al, In)
atom, i.e. the value of the polarization is negative. III-Ns are usually grown in either of
before mentioned directions, i.e., perpendicularly to the {0001} basal plane. In this way,
the abrupt change of the polarization at the interface of a heterostructure can be exploited
in device operation. Since polarization is a bulk property [40], the polarity of the crystal
does not depend on the surface layer, i.e. the termination, but solely on the direction of the
crystal structure.
Figure 2.2: Ga- and N- face of GaN. AlN and InN show similar structure. The c-axis is the polar axis in
wurtzite crystals. The group III (Al, Ga, In) and N atoms along this bond form the dipoles, which are the
basis for the macroscopic polarization. Taken from [41]
Large band-gap in GaN and AlN leads to high breakdown electric field in these materials,
3.3 MV/cm in GaN and 11.7 MV/cm in AlN [42]. These are very high fields compared to
the 0.3MV/cm in Si. Combined with high thermal conductivities of these materials, it
makes them suitable for high-power and high-temperature applications, and due to their
high saturation velocity, they can operate at high frequencies. The fact that the III-Ns are
direct band-gap semiconductors makes them a good candidate for optical applications as
well. The values of some parameters determining the electronic properties of III-Ns are
summarized in Table 1.1.
While the values of band-gap energies of GaN (3.42 eV) [43] and AlN (6.13 eV) [43] are
well established and variation in the literature is minor, there is still a considerable
2.1 Physical Properties 10
disagreement over the band-gap energy of InN (0.7 eV – 1.9 eV) [44]. Nitride ternaries and
quaternaries form a continuous range of band-gap energies Eg in between that allows for
precise band-gap engineering. The interpolation of Eg is in general not linear; it is well
approximated using a parabolic model employing a so-called bowing parameter b. For an
arbitrary parameter p, the model is expressed as
(2.1) ( ) ( ) ( ) ( ) ( )
In the discussed case, the parameter p is the band-gap energy. The band-gap bowing
parameter b takes the following values: -0.8 eV for AlGaN, -3.4 eV for AlInN and -1.4 eV
for GaInN [44]. The bowing parameters for In containing alloys assume lower values of
InN band-gap energies.
a) Reference [45]
The elastic properties of III-Ns are crucial for calculating the piezoelectric polarization.
The magnitude of this polarization makes these materials unique for employing in
electronic devices. Moreover, the areas strained extensively during the device operation are
prone to form defects and hence cause device degradation.
The elastic properties of a material describe the relationship between external forces and
internal deformations. The external forces can be described by the stress tensor σij, applied
on the crystal and related to the resulting deformation, described by the strain tensor εij.
The first index indicates the direction of stress / strain, the second one indicates the
direction that is perpendicular to the surface on which the force acts in the case of stress or
that is deformed in the case of strain. The relation between these two tensors is given by
2.1 Physical Properties 11
(2.2) or
where C is the elastic constants (stiffness) tensor and S (= C-1) is the elastic compliance
tensor. Only the symmetrical part of the stress tensor σij deforms the crystal, the non-
symmetrical part rotates it. Similarly, the strain tensor εij can be separated into a
symmetrical part representing the deformation and the rotation part. Since body torques
have no effect on polarization or defect formation, we are interested in the symmetrical
part of the tensors only, and therefore will assume that σij = σji and εij = εji. From this
symmetry follows the symmetry of the elastic tensors Cijkl = Cijlk = Cjikl = Cjilk (true for S
as well) [46, p. 132]. This reduces the number of independent components as well as
allows for reducing the number of indices by rewriting these tensors to matrices using the
Voigt notation given in Table 2.2. If a tensor is symmetrical in two of its indices, we can
use one index instead and call it the matrix notation, because the new created mathematical
object is no more a tensor, merely a matrix. This index will now run from 1 to 6.
Table 2.2: Voigt notation for reducing the number of indices of a property that is symmetrical in two of its
indices.
Matrix notation 1 2 3 4 5 6
The matrix of the elastic constants C (which is the same as that of the elastic compliances
S) is fully determined by the crystal class of the material and in wurtzites has the form
There are five (C11, C12, C13, C33, C44) independent elastic constants in wurtzites and they
are given in Table 2.3 on page 15.
2.2 Polarization in III-Ns 12
(2.3) P n p p / 0
The most of III-V semiconductors crystallize in either cubic zinc-blende (crystal class
̅ ) or hexagonal wurtzite (crystal class 6mm ) structures. Both structures meet the
2.2 Polarization in III-Ns 13
For small strains, the polarization depends on strain linearly. If there is a non-zero
polarization at zero strain, we call this polarization spontaneous. There is no other
difference between piezoelectric and spontaneous polarization. The total polarization is a
sum of both types of polarization, P = Psp + Ppz.
2.2 Polarization in III-Ns 14
Figure 2.3: The relationships between electrical and mechanical properties of a crystal. The names of the
variables (ovals) and the properties (squares) and their corresponding symbols that will be used in this work
are shown. The rank of the tensors representing the corresponding variables and properties can be deduced
from the number of their indices. Adapted from [46].
Figure 2.3 shows all variables and properties associated with piezoelectricity in crystals
and the relationships between them. Let us concentrate on the direct piezoelectric effect
(left side of the diagram) for a moment and not take the converse effect into account. This
is, in fact, the standard approach in calculating the fixed charge at the heterostructure
interface and determining the electron sheet density thus created [48]. From the diagram
we can see that, the piezoelectric polarization can be expressed, in various ways, e.g., as
[46]
(2.4) Pi pz d ijk jk d ijk C jklm lm eilm lm eilm Slmjk jk
jk jk lm lm lm lm
This set of formulae, as well as the diagram, gives (besides the expression for the
polarization) the relationships between other variables and material properties, e.g.
2.2 Polarization in III-Ns 15
The variables depend on external conditions (electric field E, stress σ, and the resulting
strain ε and piezoelectric polarization Ppz) and therefore can take different forms.
On the other hand, the forms of the tensors representing internal properties of a crystal
(piezoelectric constants e and moduli d, and the already mentioned elastic constants C and
compliance tensor S) are fully determined by the crystal class of the material. We have
discussed symmetries of the elastic tensors in Section 2.1.3. Similar considerations can be
applied to the piezoelectric tensors. The body torques do not produce electric polarization
and the electric field can distort a crystal, but does not cause a body to rotate. This fact can
be expressed assuming that both the piezoelectric constants e and moduli d are
symmetrical in their second and third indices, i.e. eijk = eikj (the same being true for d).
Similarly to the elastic tensors C and S, e and d can be rewritten in a similar manner, using
the Voigt notation defined in Table 2.2. The form of the matrix of the piezoelectric
constants e, which is the same as the one for the piezoelectric moduli d, for a wurtzite
crystal structure, is as follows
0 0 0 0 e15 0
e 0 0 0 e15 0 0 .
e 0
31 e31 e33 0 0
Only five piezoelectric constants are non-zero in a wurtzite, and only three (e31, e33, e15) of
them are independent. Table 2.3 shows the values of the properties that determine
piezoelectric and mechanical behaviour of III-Nitrides. There is disagreement over the
values of the elastic and piezoelectric constants of III-Ns in literature. The table lists some
of the most cited values. Theoretical and experimental results for parameters d and C are
discussed in an overview paper [44], and their recommended values are reported. The
theoretical calculations of Bernardini et al. [45,47,49] provide other set of values. Their
elastic constants are usually lower than those given by other authors. Moreover, the
constants C, e and d are not arbitrary; they are connected via the equation e d C .
Unfortunately, the reported values do not meet this condition, even if given by one set of
authors.
2.3 Heterostructure and 2DEG 17
Figure 2.4: Spontaneous polarization of III-N binary and ternary alloys. The thin lines show linear Vegard-
like interpolation, thick lines show the approximation to second order in disordered ternary nitride alloys,
calculated using the parabolic model (2.1) and the previously mentioned bowing parameters [53]. The
higher is the mismatch between the lattice constants of the parent binaries, the higher is the spontaneous
polarization bowing.
Linear interpolations between two binary compounds of lattice [39] and piezoelectric and
elastic [54] constants are assumed for ternary alloys. Piezoelectric moduli depend on
piezoelectric and elastic constants, and hence they depend nonlinearly on alloy
composition. However, due to nonlinear dependence of the internal parameter u on alloy
composition, the spontaneous polarization is a nonlinear function of composition as well
[55].
The Psp of III-N alloys can be expressed using the parabolic model (2.1) with the following
bowing parameters: 0.019 C/m2 for AlGaN, 0.038 C/m2 for InGaN, and 0.071 C/m2 for
AlInN [53].
Figure 2.4 compares linear and second order interpolation of the spontaneous polarization
of III-N ternaries. It is clear that the deviation is smallest for AlGaN. For In containing
ternaries, the discrepancy can be up to 40%. It has to be noted that both the linear and the
parabolic relationships are only approximate expressions and not precise formulas.
One of the semiconductors forming a heterostructure will have wider band-gap than the
other. In the case of an AlxGa1-xN/GaN heterostructure, AlxGa1-xN is the wide band-gap
semiconductor and GaN is the narrow band-gap semiconductor. At an interface of the two
semiconductors, there will be a band-gap discontinuity ΔEg, given by the differences of the
band-gap energies of the two materials. The band-gap discontinuity can be separated to the
conduction band offset (CBO) ΔEc and the valence band offset (VBO) ΔEv,
E g Ec Ev . Figure 2.5 shows an energy band diagram of a heterostructure
interface. In general, in the presence of doping in either of the two semiconductors, the
bands will bend but in III-Ns, a high electron density is created in the device thanks to the
high polarization fields and hence it is not necessary to dope the semiconductor. Therefore,
we will further assume no doping in the heterostructure.
2.3 Heterostructure and 2DEG 19
Figure 2.5: A heterojunction of two different semiconductors, which have different band-gap energies Eg
and affinities χ. The difference of the affinities determines the conduction band discontinuity ΔE c and that
together with the band-gap energy difference determines the valence band discontinuity ΔE v. In a
heterostructure, the electrons and holes see a different barrier height. Doping of either of the
semiconductors would result in band bending. However, in III-N based devices, the channel electrons are
provided by the polarization difference of the two semiconductors, which is sometimes referred to as
polarization doping. Hence there is no need for doping.
There are two models in literature to calculate the two offsets in III-Ns at a particular
fraction x from ΔEg. One splits the discontinuity evenly for all fractions x of a binary
compound in the ternary, e.g. Ec 0.68E g [56], or VBO is interpolated linearly [57].
In the latter one, all of the band-gap bowing is limited to the conduction band. A further
complication arises from the non-equivalence of the ( 0001 )- and ( 0001 )- face, which
results in a dependence of the values of the offsets on the direction of the junction [58].
E.g., for AlN/GaN (0001) ΔEv = -0.2 eV, while for GaN/AlN (0001) ΔEv = 0.85eV [59].
As will be shown in subsection 2.3.3, the difference in polarization of the materials will
produce bound charge at the interface. In the case of AlGaN/GaN heterostructure grown in
the (0001) direction, there will be a positive charge at the interface and negative at the
heterostructure surface. Figure 2.6 shows how this affects the conduction band; the effect
on the valence band is the same.
2.3 Heterostructure and 2DEG 20
Figure 2.6: The impact of interface and surface charges on the conduction band of a heterostructure. The
thin solid line represents the conduction band without any bound charges. However, the difference in
polarization of the two semiconductors induces a positive bound charge at the heterointerface. This pulls
the conduction band down evenly in the thin barrier layer (thick dashed line, which coincides with the thick
solid line in the narrow band-gap semiconductor). This charge attracts electrons on the side of the narrow
band-gap semiconductor. These electrons screen the electric field and, as a consequence, bend the
conduction band upwards. Finally, the polarization field induces a negative bound charge at the surface,
which pulls up the conduction band, as shown by the thick solid line.
Nitride devices are based on exploiting the difference in polarization between two or more
layers of Nitride alloys in creating a bound sheet charge at their interfaces. For the sake of
clarity, let us consider the simplest heterostructure, consisting of only two layers grown in
the Ga-face direction. Usually it is GaN bulk with a ternary nitride alloy on top of it.
Without the loss of generality, we can presume it to be AlxGa1-xN. Let us now investigate
the magnitude of this charge as a function of the alloy composition x. Naturally, all III-N
compounds that are used in electronics, namely AlN, GaN, and InN, have different lattice
constants in an unstrained condition. When, for instance, an AlGaN layer is grown on top
of GaN, its lattice has to be strained in order to match the underlying material.
For the purpose of further analysis, we label the c-axis and the {0001} plane of the crystal
as the z-axis and the xy-plane, respectively. As mentioned before, the direction of growth
of III-Ns employed in devices is parallel (or antiparallel) to the c-axis of the crystal, which
is normal to the {0001} basal plane. Therefore, the lattice constant of the two layers that
has to match in both of them is the side of the hexagonal base a. The six-fold rotational
2.3 Heterostructure and 2DEG 21
symmetry along the c-axis of a wurtzite structure compels the strain in x and y direction to
be the same. The strain in the basal plane can be then expressed as
a a0
(2.5) 1
a0
where a0 and a are the lattice constants of an unstrained (relaxed) and a strained structure,
respectively. In HEMT devices, the GaN layer is usually several orders of magnitude
thicker than the AlGaN layer, which is only several nanometers thick. Therefore, it is a
plausible assumption that the GaN layer will be fully relaxed, and hence show no
piezoelectric polarization, while the AlGaN layer will be strained. In this constellation, a
will be the lattice constant of an unstrained GaN (to which the AlGaN crystal structure has
to fit / stretch) and a0 the lattice constant of an unstrained AlGaN and therefore
a 0 a x
(2.6) 1
a x
where a(x) is the lattice constant of an unstrained AlxGa1-xN layer and x the Al fraction in
this layer.
The forces during the epitaxial growth of the top layer act in the xy-plane. There is no force
acting in the z-direction and there are no shear stresses or strains. The relation between
piezoelectric polarization Ppz and strain ε in formula (2.4), the form of the piezoelectric
constants e matrix, and the equality of strain in x and y direction (ε1 = ε3), tell us that in the
absence of shear strains the only non-vanishing component of the polarization vector will
be the z-component taking the following form
From the relation between strain ε and stress σ in formula (2.4), the form of the elastic
stiffness C matrix, and from the assumption of no force applied in the z-direction (σ3 = 0),
we obtain the relation between the strain along the polar axis (in the growth direction) ε3
and in the basal plane ε1 as
2.3 Heterostructure and 2DEG 22
C13
(2.8) 3 2 1
C33
Combining the previous formulas gives the expression of the piezoelectric polarization in
the strained AlxGa1-xN layer as a function of the Al fraction x (via strain in the x direction
ε1) of this layer in two alternative forms
C C 2
(2.9) P3pz 2 e31 e33 13 1 2d 31 C11 C12 2 13 1
C33 C33
Figure 2.7: The lattice constants (in this oversimplified drawing, the atoms that have to form bonds are
represented by red dots) of an unstrained AlGaN are different from that of an unstrained GaN (left). Yet,
when AlGaN grows on the top of GaN, the lattice constant a, defining the dimensions of the basal hexagon,
has to match the lattice constant of GaN. This exerts a strong stress on the grown layer in the xy-plane. This
stress causes strain (right) in both of the basal plane (ε1) and the growth (ε3) directions, described by the
equations (2.6) and (2.8), respectively. The dashed line represents an unstrained piece of AlGaN. In
piezoelectric materials, the strain results in polarization field Ppz. In this figure, the Ga-face growth is
assumed. In the case of the N-face growth, the orientation of the polarization would be opposite. a(0) is the
lattice constant of an unstrained GaN or epitaxially strained AlGaN, a(x) the lattice constant of the
unstrained AlxGa1-xN. The structure in the middle represents the orientation of the wurtzite structure in the
heterostructure.
Three things shall be noted here. First, as a by-product of the equation (2.8), it is possible
to express the lateral stress σ1 from (2.4) for epitaxially grown III-N layer as a function of
the lateral strain ε1
2.3 Heterostructure and 2DEG 23
C132
(2.10) 1 C11 C12 2 1
C33
We would like to illustrate some typical values of the lateral stress and strain in III-Ns. In
the case of Al0.28Ga0.72N/GaN heterostructure, the barrier layer grows with a lateral strain
ε1 ≈ 8×10-3 and stress σ1 ≈ 3.6GPa.
Second, the constants that enter the last two formulas, in all the three III-Ns employed in
electronic devices, have such values that the dependence of the piezoelectric polarization
C
on the strain is negative, i.e., e31 e33 13 0 for all AlxInyGa1-x-yN, where 0 x 1
C33
and 0 y 1 x . From this follows that tensile (compressive) strain, i.e., ε1 > 0 (ε1 < 0)
results in negative (positive) piezoelectric polarization, i.e., the piezoelectric polarization
vector points towards the N-face (group III-face) and hence is parallel (antiparallel) with
the spontaneous polarization and acts to increase (reduce) the polarization. The lattice
constants of AlN (InN) are smaller (larger) than that of GaN. Therefore, AlGaN grows
with a tensile strain on top of GaN, while InGaN grows with a compressive strain. The
type of strain in InAlN on top of GaN depends on the composition. The sign of the
piezoelectric constants is the same as in II-VI compounds and opposite to other III-Vs. The
values are an order of magnitude larger than in GaAs based crystals [60].
Third, the piezoelectric constants e are calculated in the equilibrium [61] and do not
describe the structure under strain. Therefore, the linear dependence given by the equation
(2.9) holds only for small strains. For typical strain values in usual III-N structures, the
nonlinearity of piezoelectric polarization in the binaries can be reproduced by a second-
order polynomial [55]:
pz
PAlN 1.8081 5.62412
(2.11) for ε1 < 0
pz
PAlN 1.8081 7.88812
for ε1 > 0
2.3 Heterostructure and 2DEG 24
pz
PGaN 0.9181 9.54112
pz
PInN 1.3731 7.55912
where ε1 is the strain of the binary compound in the basal plane, expressed by the equation
(2.6). Unlike the spontaneous polarization, the piezoelectric polarization is independent of
microscopic structure [61]. From that follows that the Vegard‟s law holds and hence the
Ppz of a ternary alloy can be calculated as
(2.12) PApzB pz
xPAN x 1 x PBN
pz
x
x 1 x N
Figure 2.8 shows piezoelectric polarization, Ppz, of III-Nitrides grown on GaN plotted
against their lattice constants a, as well as comparison between the linear and non-linear
response model to the epitaxial strain. The impact on the calculation of polarization of
AlGaN is not large (see the inset). On the other hand, for alloys with high In content, due
to high lattice mismatch and hence high built-in strain, the impact of non-linearity is more
pronounced.
The prediction for Ppz assumes pseudomorphic growth of a III-N alloy on a relaxed buffer
layer (e.g. AlGaN on GaN). That means that the upper layer (AlGaN) grows with the
lattice constant a of the buffer (GaN) and hence is fully strained. However, if the mismatch
between the lattice constants a of the two layers in a relaxed state is above a certain
threshold, the upper layer (e.g. AlGaN with a high content of Al) starts to grow partially
relaxed and with the lattice constant closer to the relaxed value. For a very high content of
Al, the AlGaN layer grows fully relaxed. For thickness of the AlGaN layer of
approximately 30nm, the degree of relaxation can be approximated by [48]
2.3 Heterostructure and 2DEG 25
0 0 x 0.38
(2.13) r x 3.5 x 1.33 0.38 x 0.67
1 0.67 x 1
The piezoelectric polarization, calculated by either the linear (2.9) or non-linear [using the
equations (2.11) and (2.12)] model, has then to be multiplied by a factor of 1 r x to
obtain a realistic prediction.
Figure 2.8: Piezoelectric polarization of III-N ternaries epitaxially grown on GaN. Dashed lines with
crosses represent linear piezoelectric response to strain, given by the equation (2.9). The spacing between
the crosses represents 10% increase in a binary alloy fraction. The solid lines represent non-linear response
to strain in binary compounds and then linearly interpolated for the ternaries. The inset shows the impact of
non-linearity of piezoelectric response to strain in the case of AlGaN grown on GaN. For Al fraction up to
65%, the error is lower than 3%.
(2.14) P dS Q
b
P b
2.3 Heterostructure and 2DEG 26
where Qb is the total bound charge enclosed by the surface Σ, ρb is the bound charge
density. From the assumption of growth in the direction perpendicular to plane {0001}
follows that the discontinuity will be in the z direction. From the absence of shear strain
during the epitaxial growth of the barrier and the form of the piezoelectric tensor e follows
that the only non-vanishing component of the polarization vector will be the z component.
To calculate the bound charge at the interface we will use the integral form of (2.14) and
construct a closed surface Σ composed of two symmetrical surfaces S just above and below
the interface and an infinitesimal surface ς connecting these two surfaces, perpendicular to
them and the interface, as shown in Figure 2.9. The bound sheet charge σb can be then
expressed in terms of the charge enclosed by surface Σ
Qb 1 1
(2.15) b P.dS PL Pu S PL Pu
S S S
where PL is the polarization in the lower layer and Pu in the upper. The polarization vector
is perpendicular to the normal of the surface ς, hence P.dS 0 and the polarization at
that surface does not contribute to the value of the integral in the equation (2.15). This
makes the third step in the equation above possible.
Figure 2.9: The closed surface Σ around an interface with abrupt change of polarization, e.g. AlGaN/GaN
interface, is used to calculate the bound sheet charge density by the equation (2.15). The polarization in the
figure is drawn in the negative direction, as it is true for all tensile and moderate compressive strains (with
no or low content of In) in III-Ns. So, if the magnitude of Pu is larger than that of PL, e.g., in the Ga-face
grown AlGaN on GaN, the bound charge will be positive.
sp pz sp pz
(2.16) b PGaN PAlGaN PGaN PGaN PAlGaN PAlGaN
P sp 0 P pz 0 P sp x P pz x
where the argument x is the aluminium fraction of AlxGa1-xN. Since the GaN layer is
usually several orders of magnitude thicker than AlGaN, it is considered to be relaxed and
hence without piezoelectric polarization, Ppz(0) = 0. The values in this equation are z-
components of P; therefore, all non-zero terms are negative. Since, according to Table 2.3
or Figure 2.4, the magnitude of the spontaneous polarization is higher in AlGaN than in
GaN, there will be a positive bound charge at the AlGaN/GaN interface. However, there is
usually a material that lacks any polarization (either air or a passivation layer) at the
AlGaN surface of the device. Hence, Pu in the equation (2.15) will be zero, and since PL is
negative, there will be a negative bound charge at the surface of the device. The situation
will be reversed in N-face heterostructure, in which there will be a negative charge at the
interface (which will attract holes) and positive at the surface. In theory, there should be
some polarization charge at the bottom GaN interface as well, but the induced electric field
is assumed to be negligible due to screening by impurities, defects and traps in the GaN
layer [56].
Figure 2.11 a) shows the values of the bound sheet charge that appears at the interface due
to discontinuity in the polarization, and the contribution of the two types of polarization, in
an AlGaN/GaN heterostructure. For low Al concentrations x, the contribution of
spontaneous and piezoelectric polarization is of similar magnitude. For higher x, the
AlGaN barrier is more and more relaxed and less strained, and hence the piezoelectric
polarization of that layer drops down. For very high Al concentrations, the whole bound
sheet charge is only due to the difference in spontaneous polarizations between the two
layers.
2.3 Heterostructure and 2DEG 28
2.3.4 2DEG
Figure 2.10: Vertical cross-section of the conduction band in the heterostructure. The dotted line is the
position of the Fermi level in the semiconductor. Φb is a barrier height. In the case of a heterostructure or a
HEMT far from the gate the barrier height is determined by the surface sheet charge. Under the gate of a
HEMT, it is determined by the Schottky barrier, modified by the applied gate voltage. Δ is the penetration
of the conduction band edge below the Fermi level at the AlGaN/GaN interface, ΔEC is the conduction
band offset, E0, is the lowest subband level of the 2DEG. The labels correspond to the ones used in
equations (2.17) and (2.18). Adapted from [62].
The large polarization difference at the heterostructure interface produces large bound
charge, which in turn gives rise to a high electron sheet density, ns. The electron density is
affected by the barrier height Φb [63], and the electron sheet density is calculated as [64]
b x 0 x
(2.17) n s x, d eb x x Ec x
e e2d
where Δ is the penetration of the conduction band edge below the Fermi level at the
ABN/GaN interface (A and B stand either for Al, In or Ga), ΔEC is the conduction band
offset, ε0 is the dielectric constant of vacuum, ε is the relative dielectric constant of the
barrier layer. Δ is calculated using the expression
23
9e 2 n x,d 2
(2.18) x s n s x, d
8 8m x
* m*
0
where the first term, in the Figure 2.10 labelled as E0, is the lowest subband level of the
2DEG with the effective electron mass m* ≈ 0.0228 me [65]. It follows from the equations
(2.17) and (2.18) that the formula to calculate electron sheet density ns is itself dependent
2.4 AlGaN/GaN HEMTs 29
on ns, thus the calculation has to be carried out self-consistently. Figure 2.11 b) shows the
2DEG concentration as a function of barrier thickness for various Al concentrations.
a) b)
Figure 2.11: a) Contributions of spontaneous and piezoelectric polarizations to the creation of the bound
sheet charge at the AlGaN/GaN interface. The dotted line represents piezoelectric polarization in fully
strained layer. However, the AlGaN layer has been found to relax [48] for higher strains. If we take the
relaxation of the layer to account, as given by (2.13), the piezoelectric polarization follows the dependence
that is shown as the thick dashed line. The difference in spontaneous polarizations between AlGaN and GaN
layers is the only contribution for high Al concentrations. For lower concentrations, it is similar to the
piezoelectric contribution. b) The dotted lines represent the bound charge at the interface (calculated in the
subfigure a as P_total) for various Al concentrations. As a consequence of the bound charge, 2DEG forms
in the GaN channel, as given by the equation (2.17). The solid lines represent the electron density as a
function of the AlGaN barrier thickness. For every Al concentration, there is a critical barrier thickness under
which no 2DEG is formed.
2.4.1 Introduction
A transistor is a semiconductor device with three or more terminals. Its operation is based
on controlling a signal at one pair of terminals by a signal applied at another pair of
terminals. The main functions of a transistor are to amplify or to switch electronic signals.
In a Field Effect Transistor (FET), the current flows between the source and drain
terminals (ohmic contacts) through a channel. The channel conductance is modulated by an
electric field perpendicular to the surface produced by the voltage applied between the
source and gate. In this type of transistors, only the majority carrier is involved in its
operation. The gate can be separated from the channel by an insulator (as in a MOSFET),
2.4 AlGaN/GaN HEMTs 30
can form a pn junction (JFET), or a Schottky barrier junction with the channel [MEtal
Semiconductor FET (MESFET)]. A modification of the MESFETs is the High Electron
Mobility Transistor (HEMT), which utilizes a heterostructure to create a potential well
perpendicular to the heterointerface. The electrons that are confined to this potential well
are free to move parallel to the interface, forming a 2DEG. The conventional HEMTs are
GaAs based. Arsenide III-Vs do not possess spontaneous polarization and the piezoelectric
constants are an order of magnitude lower than those of nitrides. Therefore, undoped III-V
arsenide heterostructures cannot induce high 2DEG. Even with doping, the conventional
GaAs based HEMTs can achieve 2DEG density of approximately 2 × 1012 cm-2. However,
intentional doping is not necessary for GaN based devices, since, due to the high
polarization, the 2DEG densities are already on the order of 1013 cm-2. Moreover, doping
could reduce the electron mobility via scattering.
Figure 2.12: Schematic structure of a HEMT device. The figure is not to scale. In a real transistor, the
length of the device is much larger than the thickness of the AlGaN barrier. The 2-dimensional electron gas
(2DEG) is in the potential well in the GaN layer, which is the lower band-gap semiconductor in this
heterostructure, near the heterostructure interface. The 2DEG creates the channel, which leads current in
the device. The current flows between the ohmic contacts, the source and the drain, and is modulated by the
voltage applied at the gate, which is a Schottky contact.
Figure 2.12 shows a schematic view of a GaN-based HEMT device. The gate is placed
asymmetrically. At high drain voltages, the electric field between the gate and drain can be
very high, having a peak at the drain edge of the gate. The gate is shifted away from the
drain to reduce this field and hence increase the breakdown voltage of the device.
However, increasing the distance between the gate and drain has a negative impact on the
performance of high frequency devices, especially reducing the cut-off frequency.
2.4 AlGaN/GaN HEMTs 31
c) closed, VG < VT
Figure 2.13: Closing the channel with increasing the negative gate voltage VG. Positive (negative) voltage
applied on the gate lowers (increases) the Schottky barrier Φb. The solid line represents the conduction
band and the dotted line represents the Fermi level. EF,m is a) A fully open channel where the applied VG is
positive, but this is not necessary. b) Conduction band of a HEMT with negative applied VG, but the
channel is still open. c) The negative VG is larger than the threshold voltage VT and the channel is closed.
b x 0 x
(2.19) ns x, d eb x VG x Ec x
e e2d
Figure 2.13 shows the band structure of the device under the gate for various gate voltages
and hence various regimes of operation. There is a high 2DEG density in the channel in the
saturation regime; the density gets smaller with the gate voltage approaching the threshold
voltage VT in the linear regime; and finally, there are virtually no free electrons to carry
current when the channel is closed (as the negative VG exceeds VT).
The conducting channel (under the heterointerface, between the source and drain
terminals) can be viewed as a resistance. For small drain-source voltage VD, the drain
current ID is approximately linear. When a negative voltage is applied to the gate, the
2.4 AlGaN/GaN HEMTs 32
electrons are partially depleted from the channel and its resistance increases. As the
negative gate voltage VG is increased, a threshold voltage VT is reached. At the threshold,
the channel is closed, i.e., completely depleted of electrons, and the ID drops to zero. This
condition is called pinch-off. The evolution of the conduction band with respect to
changing VG is shown in Figure 2.14 a). Such a transistor, which is switched on (the
channel is conducting) at zero VG, and requires a negative gate voltage to shut down the
current, is called depletion mode or normally on transistor. A device, which is off at zero
VG, and requires a positive gate voltage to switch the device on, is called enhancement
mode or normally off transistor. If VD is increased at a fixed VG, the drain edge of the gate
begins to be reversely biased in respect to the channel, which causes electron depletion in
that region. Now, the channel resistance becomes position dependent and since the current
has to remain constant, all the additional voltage drops in the region with high resistance,
near the drain edge of the gate. The electric field in the source region will not increase
anymore and the current will saturate. The potential across the channel is shown in
Figure 2.14 b) and the electric field close to the source in its inset.
a)
b)
Figure 2.14: The conduction band across the channel. Gradual closing of the channel due to reverse biasing
the gate at a constant drain voltage is shown in the subfigure a). The x-component of the electric field close
to the source is shown in the inset. As shown in the subfigure b), for small drain voltages, the voltage drop
is spread across the whole channel. However, as the drain edge of the gate becomes depleted, any
additional voltage applied drops in that region and the electric field near the source does not increase
anymore. This is the saturation region. Again, the inset shows the x-component of the electric field near the
source. The drain current is determined by this electric field. The position of the contacts is indicated at the
bottom of each subfigure.
2.4 AlGaN/GaN HEMTs 33
The term traps refers to energy states in the band-gap of a semiconductor. The origin of
traps can be a consequence of several factors, e.g. crystal defects, dislocations, or the
presence of impurities. These trap states may be empty or occupied by electrons, which has
an impact on the charge they carry. Trap states in the upper part (above the neutral level) of
the band gap (closer to the conduction band) are acceptor-like, neutral when empty and
negatively charged when occupied. Trap states in the lower part (below the neutral level)
of the band gap (closer to the valence band) are donor-like, positively charged when empty
and neutral when occupied. Traps at the interface or the surface of a device play an
important role in the device operation and performance.
One of the conclusions of Section 2.3.3 was that there is large negative bound charge at the
surface of the device. The negative charge would repel the electrons away from the
interface and deplete the channel. Moreover, there is a question, what is the origin of the
electrons in 2DEG. It was suggested that, after the growth, during the cooling process, free
electrons would compensate the polarization-induced charge [64]. As a different solution
to both problems, nowadays widely accepted, it was suggested that surface donor-like traps
could be the source of both the channel electrons and the positive charge screening the
large negative polarization-induced bound charge [66,67]. For low AlGaN barrier
thickness, the surface trap level is below the Fermi level, the traps are occupied and hence
neutral. At a critical barrier thickness, the surface traps reach the Fermi level and the
electrons from these traps are driven into the channel by the strong polarization-induced
electric field in AlGaN [56]. The band diagram, with the energy level of the surface traps,
with varying barrier thickness is shown in Figure 2.15 a). As the donor-like traps are
emptied, they become positively charged and in effect they reduce (passivate) the negative
bound charge. In the absence of holes, the energy of these traps has been theoretically
predicted to be ~1.65 eV in Al0.35Ga0.65N [67] and, by fitting simulations to experimental
data, determined to be in the range of 1.42 eV [66] to 1.85 eV [56] below the conduction
band, both in Al0.27Ga0.73N layer.
2.4 AlGaN/GaN HEMTs 34
At large negative gate voltage, electrons from gate may leak to the trap states in the
ungated surfaces and create a „virtual gate‟ and modulate the depletion region [68]. The
corresponding charge distribution in the device is shown in Figure 2.15 b). In pulsed
operation, the gate voltage changes abruptly and since the response of the trapped electrons
is not immediate, it leads to RF drain current collapse phenomenon. The transient time
constants depend on the energy level of the traps [69]. The surface donor-like traps used to
explain the origin of 2DEG can explain RF current collapse with time constants on the
order of seconds [70], as observed experimentally [71]. However, transients with shorter
time constants (10-100 μs) can be explained by the existence of surface donors with energy
level 0.3 eV [70] resp. 0.25 eV [72] above the valence band.
a) b)
Figure 2.15: The effect of surface traps on 2DEG creation, reduction of the negative surface bound charge
and current collapse. a) The polarization charge induces large electric field in the AlGaN layer. For thin
AlGaN barrier, the trap energy level is below the Fermi level and the states are filled. As the barrier reaches a
critical thickness, the trap level hits the Fermi level and the traps start to empty and become positively
charged and, due to the strong electric field, the electrons transfer to the channel [66,56]. The diagram is only
schematic, in reality, the bands in GaN change with the barrier thickness. For a thin barrier, all electrons are
in the traps and not in the channel; hence the conduction band will be above the Fermi level and will not
bend. b) Large negative VG bias will induce high electric field close to the gate and the electrons form the
gate leak to the empty surface states, create a “virtual gate” [68] and deplete the channel. In pulsed operation,
after the abrupt change of VG, it takes some time to remove the trapped electrons and for that time the channel
remains partially depleted. This reduces the drain current and hence the expected output power. This
phenomenon is called RF current collapse.
2.4 AlGaN/GaN HEMTs 35
2.4.3 Substrates
The most widely used substrates in the GaN technology are Si, sapphire (Al2O3), and SiC.
Recently, diamond and GaN have been employed as substrates as well. The advantage of
the diamond is its high thermal conductivity, while the using of GaN reduces the density of
impurities. One of the main tasks of a substrate is to conduct and dissipate the heat
generated during device operation. This process is governed by the heat flow equation [73]
(2.20) c
T
t
T H
where c is the heat capacity, κ is the thermal conductivity and H is the heat generation rate,
Table 2.4: Material parameters that determine heat dissipation by the substrate. a) Reference [28], b)
Reference [74].
Thermal
Heat Capacity
Substrate Conductivity Density (g.cm-3)
(J.g-1.K-1)
(W.cm-1.K-1)
0.35 a 0.77 a 3.98 a
Al2O3
0.42 b
Si (111) 1.5 b
3.6 a 0.66 a 3.21 a
4H-SiC
3.3 b
1.6 a 0.49 a 6.1 a
GaN
1.7 b
2.4.3.1 SiC
Lattice mismatch between SiC and GaN is 4%. Thanks to its very good thermal
conductivity, it is the most attractive substrate. Layers of GaN grown on this substrate
exhibit an excellent crystallographic quality: the density of dislocations is under 3x10 8 cm-
2
, thanks to a nucleation layer of AlN, which ensures a smooth transition between the
crystal structure of SiC and GaN. Of all substrates, it is the preferred one for high
frequency applications. Unfortunately, SiC is very expensive.
2.4 AlGaN/GaN HEMTs 36
Of the commonly employed substrates, sapphire has the largest lattice mismatch with GaN.
Depending on their relative orientation to each other, the mismatch is between 14% and
23%. The main disadvantage of this material, however, is its poor thermal conductivity.
This is a problem especially in applications, in which it is necessary to dispose of heat
effectively, and it may result in overheating the device. On the other hand, the advantage of
this substrate is that it is cheap and available in wafers with large diameters.
2.4.3.3 Si
improvement of the technology in the last decade, the figure includes also the first RF data,
measured in 1996 [12].
2.5.1 Self-Heating
As mentioned earlier, the strength of GaN-based HEMTs, due to their large band-gap, is
operation at high voltages. During high-voltage operation, high electric fields and current
densities are induced, which in turn generate large amount of heat [73]. Self-heating of the
device leads to increase of the lattice temperature, and hence to deterioration of transport
properties [93,94]. The performance in high power operation will depend on the quality of
the substrate, namely on its thermal conductivity, since it is desirable to dissipate as much
heat as possible. The thermal behaviour of GaN-based HEMTs, with emphasis on the
influence of the substrate, is extensively investigated. Using micro-Raman spectroscopy
[28] for SiC and sapphire substrates, it was shown that the increase of temperature on SiC
2.5 Key Challenges in Current GaN Technology 38
is slower than on sapphire. Moreover, on the sapphire substrate, the reached temperature is
higher. It was found that the rise and fall of the temperature after switching the device on
or off, is very fast for both substrates, below 200 ns. Therefore, apart from DC operation,
this phenomenon will affect RF or pulsed operation as well. Using electro-thermal Monte
Carlo method [74] for SiC, sapphire, Si, and GaN substrates, it was shown that SiC
provides the highest current and the lower peak in temperature and hence is the most
suitable substrate for high-power applications. Recently, it was shown that due to thermal
effects, the current saturation occurs at a lower electric field than the field at which current
saturates in the bulk GaN [95], indicating again, that the thermal effects play a
considerable role in the GaN-based HEMT operation.
In the literature, the term current collapse is used with two meanings. In a wider sense, it
means a class of phenomena that lead to drain current degradation. Historically, however,
it was defined, during the development of GaAs-based HEMTs, as a persistent yet
recoverable reduction of the dc current at a high VD [96]. Some of the other phenomena
that fall under the wider meaning of the current collapse are gate and drain lag, and DC-RF
dispersion [96]. Gate (drain) lag refers to ID transient in response to gate (drain) voltage
pulses keeping the drain (gate) voltage constant. The corresponding measurement
techniques are combined in the I-V pulsed measurement, in which both the gate and drain
voltages are pulsed at the same time from a quiescent bias. The difference between DC
operation and I-V pulsed measurement is referred to as DC-RF dispersion, also known as
“knee walkout” [97,26] due to the representation of the effect in the I-V plane, illustrated
in Figure 5.1 b). All of these effects are recoverable. The leakage current from the gate,
and carrier trapping in general, into the surface (as described in Section 2.4.2) and buffer
traps is the physical cause behind the current collapse class of the phenomena.
Device degradation refers to the unrecoverable ID degradation, which occurs after stressing
the device by a high voltage for long periods of time, typically for several hours. The effect
of the device degradation on the I-V characteristics is similar to that of the DC-RF
dispersion. After exhaustive stress experiments, in which it was established that, the device
is degraded by operation at high voltages, and not as much at high currents, following
2.6 Summary 39
explanation of the phenomenon was offered [29]. At high voltages, and hence high electric
fields, excessive stress is induced through the inverse piezoelectric effect [98]. The stress
forms lattice defects, which act as traps. Thus, the device degradation is also caused by
trapping of carriers in the surface and in the buffer. The difference between the current
collapse and the device degradation is that, the former is caused by trapping carriers in
existing traps, while the latter by creating more traps, probably of energy levels that retain
the trapped charges. The recoverable current collapse due to trapping in surface states has
been largely suppressed by the means of passivation. On the other hand, device
degradation is still a pending problem in GaN technology, and we may quote from the
ITRS again, “key challenge in GaN technology … [is the] reduction of leakage current and
understanding of failure mechanisms” [1].
2.6 Summary
This chapter discussed the properties of III-N semiconductor materials that are most
important for the operation of devices fabricated from this class of materials, such as
electrical and elastic properties, and the HEMT devices based on heterostructures from
their alloys. Special attention was given to the polarization in III-Ns, because this quantity
is much stronger in these materials than in other semiconductors and has a strong influence
on the operation of these devices. The difference in the polarization in two III-N alloys at
their interface leads to polarization induced bound charge which attracts electrons and
creates high concentration 2DEG without the need for doping. Also, the interplay between
the mechanical and electrical properties, i.e. the direct and converse piezoelectric effect,
was discussed, because this plays an important role in some phenomena affecting the III-N
based HEMTs, e.g. in device degradation. The converse piezoelectric effect is elaborated
in Chapter 4, where it directly affects the electric field induced polarization and also taken
into account in Chapter 5 in investigation of the device degradation. The principle of
HEMT operation was discussed along with related topics, such as the influence of surface
traps, the substrates used for these devices. An overview of current front end devices was
given, as well as current challenges facing the nitride-based devices. The following chapter
will exploit the principles laid out here with the final aim of calibrating two devices using a
commercial simulator, also introduced in the next chapter.
3 Simulation Methodology
This chapter describes the simulation methodology used in the thesis. All simulations are
carried out with the commercial simulation tool Sentaurus. The simulation platform is
described in section 3.1. The underlying drift diffusion approach is described in section
3.2. The scripts developed to manipulate and have better control over the Sentaurus
simulation tool are described in section 3.3. Finally, the calibration procedure is described
in section 3.4.
Sentaurus, which is the simulation platform used in this thesis, is a TCAD (Technology
Computer-Aided Design) simulation tool from Synopsys [99], which solves a system of
partial differential equations to model the electrical behaviour of semiconductor devices
[100]. Synopsys TCAD offers a set of simulation tools for process and device simulations.
In this work, the following tools were used. Sentaurus Workbench (SWB), a flexible
framework environment with advanced visualization and programmability, Sentaurus
Structure Editor (SSE), a 2D/3D device editor, Sentaurus Device (SD), a 2D/3D device
simulator, and the tools for visualising the simulation results, Inspect, used for viewing
one-dimensional functions, e.g. ID-VG, and Tecplot SV, used for viewing the distributions
of parameters in the device in 2D and 3D. A description of selected tools, by no means
exhaustive, follows.
3.1 The Simulation Platform 41
Figure 3.1 shows an example of a project, in a Projector Editor view of SWB graphical user
interface (GUI). In this specific project, SSE has one parameter associated with it (x), SD
has three parameters (Vgmin, Vgmax, Vd), and Inspect has none. However, the parameter x,
that stands for the fraction of Al in the AlxGa1-xN barrier layer, is read by both SSE and SD,
and the three parameters associated with SD are read by Inspect as well, to plot the correct
ID-VG characteristics. SD performs simulations of ID-VG in the range between Vgmin=-6V
and Vgmax=2V at ten different values of VD. The global parameters make it possible to
have general input files and still have consistency between them. Before running the
simulations, it is necessary to pre-process the files to convert the parameters to the values
specified in the project and to make necessary calculations within the command files using
a scripting language of the specific simulation tool.
3.1 The Simulation Platform 42
Figure 3.2: Simulation flow of the Sentaurus simulations platform. Only the tools that were used in this work
are mentioned. Not all input/output files are shown. Each simulation tool here corresponds to an input file.
First, using SSE, files defining the MESH and DOPING are created. Using these files and the simulation tool
SD, the device simulation (set of voltage sweeps at one of the electrodes) is performed and two types of
output files are produced. CURRENT, that records the current and voltage at all electrodes, and DATA, that
records all pre-specified simulated parameters in all mesh points, e.g. the electric field, the conduction band,
the current density, the electron density, etc. Inspect and Tecplot SV visualise the CURRENT and DATA files,
respectively.
SSE is used to define the structure of a device, including the electrodes, doping etc. [102].
In order to create the device structure, it is possible to either directly write/modify the input
file or use the GUI, which will create the input file for the user. The SSE tool will create
the actual device structure based on the input file. The output of SSE are several files, two
of which are of interest to us (Figure 3.2), a file defining the MESH and a file defining the
DOPING in the device that is matched to the mesh points defined in the MESH file. The SSE
command file has several sections. At the beginning, the user may define variables that
will be used later in the file and use simple mathematical operations on them. The
advantage of doing this is that a single variable may be used in various places in the input
file and so to change all instances, it is necessary to change the definition of that variable
only. This is true for all Sentaurus tools‟ command files. Next, using rectangles or other
polygons, the device structure is defined as regions of different materials, with names
given to each region, to be used as a reference later in the file and in the SD command file.
Doping profiles may be defined in this input file. High concentration donor doping is used
in a small region around the ohmic contacts (source and drain) of the device to emulate the
3.1 The Simulation Platform 43
Figure 3.3: HEMT structure created in SSE, showing the mesh. GaN layer is in grey colour, insulator
between the contacts is in red, AlGaN layer between these layers is not visible in this scale due to its thinness
and the density of the mesh in that region. The three dents on top of the device, represented by the green
lines, not originally shown in the SSE, are the electrical contacts; from left to right it is the source, the gate
and the drain.
metal spikes. The electrodes are placed at some of the edges of the device. Finally, the
mesh density is defined by creating windows in a shape of lines, rectangles or polygons
and specifying the mesh density at the ends of the window. At last, using all the
specifications, another Sentaurus tool Mesh is called from the SSE to generate the mesh in
the MESH file with doping concentration assigned to each mesh point in the DOPING file. A
typical mesh is shown in Figure 3.3 in an SSE window with the highest mesh density
around the gate. A detail of the gate region is shown in Figure 3.4.
Figure 3.4: Detail of the region under the gate of the HEMT device structure shown in Figure 3.3.
3.1 The Simulation Platform 44
The mesh density is an important aspect of device simulation. Increasing the density has
positive impact on the precision of the calculation of current in the device and other
parameters of interest, but also negative impact on simulation time. It is therefore
necessary to balance the need for precision and reasonable simulation time. High density of
the mesh is used in regions important for the operation of a device, i.e. where there are
large gradients of parameters such as the electric field or current density. The region of
high mesh density in the vertical direction is usually the heterointerface and the channel,
and in horizontal direction it is the gate edges, especially the drain edge. In our
simulations, the vertical mesh spacing at the AlGaN/GaN interface was around 4Å, the
horizontal mesh spacing under the gate was up to 4nm, near the drain edge limited to
approximately 1nm. Between the gate and the drain, far from both electrodes, the mesh
was sparser, in the order of 10nm. Total mesh size of the simulated devices was in the
order of tens of thousands of mesh points. In the investigation of current collapse in
Chapter 5, where we placed rapidly changing surface charge, and especially in Chapter 6,
in which we investigate trap to trap hopping of electrons along the device surface, the
region of high mesh density was extended further towards the electrodes. As mentioned
above, reducing the mesh spacing and hence increasing the number of mesh points leads to
higher precision of the simulation until a limit is reached above which only the simulation
time is increased. The optimal mesh size depends on the investigated problem. We haven‟t
encountered any mesh size related convergence problems.
Sentaurus Device is a tool to define physical models used in the simulation, the model and
material parameters and to define and run the actual simulation [103]. Two types of input
file are necessary to run a simulation using this tool. A material parameter file, one for
each material defined in the device, and a command file, with parameters relating to the
whole device.
The command file is divided into several sections governing different aspects of the
simulation. The section Electrode defines the types of used electrodes (ohmic, Schottky
contact, etc.), the initial voltage applied to the electrodes and parameters depending on the
type of the contact, e.g. work-function for Schottky contact or contact resistance for ohmic
3.2 Carrier Transport 45
contacts. The section Physics defines the physical models to be used in the device
simulation e.g. mobility models, generation-recombination models, etc. The models can be
global or specific for a region, a material, an interface or an electrode. Other quantities that
can be defined in this section are a fixed charge at a specified interface, traps in the bulk
with either uniform or Gaussian profile, etc. The section Plot specifies the quantities
calculated in the simulation to be output in the DATA file for each mesh point at specified
moments of the simulation. The section Solve specifies what is actually simulated, i.e.,
which equations are solved, e.g. Poisson‟s equation (3.2), current continuity equation for
electrons and holes (3.3), series of target values to which the voltage of specified
electrodes should change, e.g. for I-VG or I-VD characteristics. Additionally, it is possible
to specify the names for the CURRENT output file for sections of the simulation and specify
when to save the DATA output files. Moreover, it is possible to save the state of a
simulation and load that state later in the command file.
The material parameters that must be specified in the input file include among others the
relative permittivity, temperature dependent band-gap, affinity, electron and hole effective
mass or density of states. Other parameters depend on the selected physical models defined
in the command file. The drift-diffusion simulation requires a mobility model, defining the
low-field mobility, and other model-dependent parameters, such as the saturation velocity.
To describe the transport of carriers in a device, we will first introduce the concept of
mobility (3.2.1), which relates the motion of the carriers in the device to the electric field.
Then, we introduce the fundamental equations that govern the transport of carriers (3.2.2),
i.e., the Poisson‟s equation, which is used to calculate the electrostatic potential Ψ self-
consistently with the electron and hole concentrations, n and p, and the current continuity
equations for electrons and holes.
To close the circle and make the solution self-consistent, it is necessary to calculate the
current densities, Jn and Jp, from the electrostatic potential Ψ and the mobile charge
concentrations n and p. This is accomplished by one of the transport models. Sentaurus
Device implements the following models: Drift-diffusion (DD), Thermodynamic (TD),
3.2 Carrier Transport 46
DD model cannot capture velocity overshoot which becomes crucial in deep submicron
devices and is not accurate in estimating the impact ionization generation rates [103]. HD
model overcomes these deficiencies of DD model at some computational cost. In our work,
we have not investigated or considered the impact ionization and the dimensions of the
devices was on the order of microns. For these reasons, we have used only the Drift-
diffusion approach and therefore, from now on we will concentrate on this model (3.2.3).
3.2.1 Mobility
(3.1) ,
where is mobility, a measure of response of the ensemble of carriers to the electric field
E, and accounts for positively/negatively charged particles, i.e., holes and electrons. The
mobility is determined by a variety of scattering mechanisms including phonon, ionised
impurities, surface roughness and other types of scattering.
(3.2) ( ),
3.2 Carrier Transport 47
where n and p are the electron and hole concentrations, ε is the dielectric constant, ψ is the
electrostatic potential, q is the elementary charge, is the concentration of ionized
donors and is the concentration of ionized acceptors.
The time dependent current continuity equations for electrons and holes are given by
(3.3) and
where R is the net electron-hole recombination rate (recombination – generation). (3.2) and
(3.3) are the basic equations for simulating the carrier transport in semiconductor devices.
In a steady state, equation (3.3) becomes
(3.4) and
Charged particles in motion give rise to electric current. The electric field sets the charge
particles into directed motion and in a semiconductor the average drift velocity of such
charged particle is given by equation (3.1). Since the current is defined as a flow of
positively charged particles, the current density vector has the direction of the drift of the
holes and in the direction opposite to the drift of electrons. Hence, using (3.1), the drift
current is given by
(3.5) ( ) ( ) ,
current, oriented in the direction of the diffusion of holes and in the direction opposite to
the diffusion of electrons, which leads to
(3.6) ( ),
In the absence of any external bias applied to the semiconductor, the combined drift and
diffusion current must be, both for the electrons and holes, equal to zero. Taking into
account the expression linking the electron concentration to EF-Ei, i.e. the Boltzmann
relationship for electrons,
(3.7) ( )
(3.8) ,
This relationship links the two constants from the expression of the drift (3.5) and diffusion
(3.6) currents and shows that they represent the same quantity and differ only by a
multiplication coefficient. In the previous equation, EF is the Fermi level energy, i.e. the
highest occupied energy level at 0K or energy level with 50% probability of being
occupied at T > 0K, Ei is the intrinsic energy level, i.e. the level of EF in an intrinsic
semiconductor, and ni is intrinsic carrier concentration, which is a function of only
temperature and band-gap. Under thermodynamic equilibrium conditions, EF for electrons
and holes is the same. For non-equilibrium conditions it is not true, but (3.7) holds, if
separate quasi-Fermi levels are assumed for electrons and holes
(3.9) ( )
( )
3.3 Scripts 49
Substituting the Einstein relationship into the drift and diffusion current and using the
definition of quasi-Fermi levels, the combined drift-diffusion current density can be
expressed as
(3.10) and .
The underlying assumptions of the drift-diffusion model include the relaxation time
approximation, equal temperatures of the carriers and the lattice and a slowly varying
electric field [104].
3.3 Scripts
In order to accomplish the tasks set out in this work, a number of scripts were developed.
The tasks can be grouped as follows. Firstly, the scripts included calculations performed
outside the scope of the Sentaurus simulator, based on the results from a previous
Sentaurus simulation and automatically read from an output file. Based on these
calculations, the input files for the next simulations were rewritten and a following
simulation performed. Secondly, the scripts included automated loops of Sentaurus
simulations to generate a family of results with varying selected parameters in order to
calibrate, i.e. find specific values of the parameters that reproduced the measurements, the
investigated physical processes. Thirdly, the scripts included automated evaluation of the
large number (thousands) of simulation results.
Since some of the subtasks were required repeatedly in different tasks, the appropriate
developed procedures were reused multiple times. These include ReadCoords that reads
the coordinates of a mesh that represents a device from the MESH file, ReadData that reads
a property from the DATA output file in all of the mesh points, WriteData that writes a file
in the format of the DATA output file with calculated parameters so that they can be easily
visualised using Tecplot SV, ReadDoping that reads the doping concentration from the
DOPING file that is used by the simulator and WriteDoping that rewrites the doping
concentration in the DOPING file. In the following we offer description of some of the
scripts used in this work.
3.3 Scripts 50
Shdop: simply puts sheet charge, converted to charge concentration, at a line with one axis
constant and between two points at the other axis, according to the values defined in the
script‟s input file. To accomplish this task, the script uses ReadDoping, to preserve any
doping already present in the device, and WriteDoping that writes the original doping with
the addition of the specified sheet doping.
Polariz: was used in the simulations of Polarization Induced Bound Charge, reported in
Chapter 4, using equation (4.8) for the bound space charge generated by the converse
piezoelectric effect and equation (4.9) for the sheet space charge at the AlGaN/GaN
interface modified by the same effect. The simulation flow is illustrated in Figure 4.5. In
this script, ReadCoords and ReadData read the MESH and DATA files from an initial
simulation to access the electric field calculated by Sentaurus and to assign the value to a
specific mesh point with its coordinates. Then, the actual calculation of several quantities
of interest, such as the lateral stress σ3 (4.4), the vertical strain ε1 (4.3), z-component of the
polarization P (4.7) and the bound charge ρb, was performed. Since Sentaurus requires the
space charge to be defined in the simulations, the sheet charge was converted to the space
charge. Finally, WriteData output the calculated quantities.
ChrgDop: Reads the DOPING file by ReadDoping and the DATA file produced either by
Polariz or StressChrg, searches for the quantity named “Charge” and writes a new DOPING
file by WriteDoping with the original doping and the new charge-as-doping combined. In
the investigation of the impact of the converse piezoelectric effect on the I-V
characteristics, Polariz and ChrgDop were repeated in a loop until the electric field in the
device converged.
CurrColl: was used to calibrate a proposed model of trapped charge at the surface
responsible for the current collapse phenomenon, in section 5.4. The script consists of a
loop, in which all predetermined values of the searched parameters are selected. Inside of
3.3 Scripts 51
this loop, Shdop is called to insert an exponential charge distribution at both sides of the
gate and a constant charge distribution under the gate. Then, SD is called to perform the
device simulation with the created doping file. Finally, the I-V characteristics is archived
for further analysis.
Extract: was used to extract and evaluate the set of the simulated I-V characteristics,
produced by CurrColl, by automatically comparing them to the experimental
measurements. First, the appropriate experimental ID/G-VD/G is read. Then, the first
simulation I-V SD output in the CURRENT format is read. In general, the values of V in
both files are different. Therefore, the next step is to calculate the simulation current at
voltages at which the current was measured. Next, the deviation is calculated as a square of
the differences between the measured and simulated current for all points and the values
are added, and the final value is recorded. Continue with reading the next simulation I-V
until all are analysed. Find the simulation with the lowest deviation. The whole procedure
is described in more detail in section 5.4.2. A modified version of this script was also used
in the assessment of the accuracy of the calibration in this chapter, in subsection 3.4.3.3,
where it was necessary to compare the measured and the simulated I-V characteristics at
bias points that were different in the two sets of data.
PF: was used to simulate the Poole-Frenkel emission and transport mechanism, suggested
to be responsible for the electrons leaked to the surface of the device causing the current
collapse mechanism, reported in Chapter 6. This procedure is described in detail in
section 6.4.
3.4 Calibration of the Simulator 52
Figure 3.5: The electron mobility in a GaN HEMT as a function of electron concentration at three different
values of the temperature [105]. The symbols are at VG = 0V.
The aim of the calibration is to reproduce the measured I-V characteristics or other
dependence in the simulation, using realistic physical models and realistic values of
physical parameters. The parameters of interest are the bound charge at the interface σi and
surface σs of the device, the band-gaps, the permittivity in both GaN and AlGaN, electron
effective mass , low-field electron mobility μn0 and saturation velocity vsat in GaN, the
contact resistance of the ohmic contacts (source and drain) Rc, and the gate metal work-
function Wf, which determines the Schottky contact barrier as , where χ is
the affinity of a semiconductor, in our case of AlGaN, since the gate is on top of this
material. The (positive) bound charge σi can be calculated using equation (2.16) and the
materials‟ electromechanical properties listed in Table 2.3. σs can be calculated from the
more general equation (2.15) and the same properties, but determining this parameter
accurately becomes more complicated. As was explained in subsection 2.4.2.1, this
(negative) charge is partially compensated by emptying the donor-like traps that become in
effect positively charged. Hence, σs becomes a fitting parameter.
assuming local relation between the velocity and the electric field. One of the most
common field dependent mobility models is the Caughey-Thomas model [106] illustrated
in Figure 3.6, expressed as
(3.11) ⁄ ,
( ( ⁄ ) )
where μn0 is the low-field mobility, vsat the saturation velocity and β is a fitting parameter
that controls how smooth is the transition between the linear and saturated region of the
dependence of velocity on the electric field, for electrons usually set to 2.
Figure 3.6: Mobility and velocity as a dependence of the electric field parallel with the carrier current in the
Caughey-Thomas mobility model [106]. The mobility in low electric field is governed by μn0 and the velocity
in high electric field by vsat, hence the names of the parameters.
Several parameters used in the calibration can be measured independently. This includes,
for example, Rc. For Ni gate, ΦSch dependence on the Al fraction was measured [107,108]
and calculated [109] and assumed to be a linear function. The relative permittivity was
reported to be 10.4 for GaN and 10.1 for AlN [48]. It is possible to estimate this parameter
for the barrier layer of AlGaN based on the fact that it determines the slope of the
3.4 Calibration of the Simulator 54
dependence of the electron sheet density on the gate voltage, given by equation (2.19), if
there is such a measurement at disposal. in GaN was experimentally measured to be
0.20 [110,111] with approximately 1% anisotropy [44]. The experimental values of μn0
obtained experimentally at 300K and used in simulations range from 1070cm2.V-1.s-1 [23]
to 2000cm2.V-1.s-1 [112] and the values of vsat range from 1×107cm.s-1 [23] to 2.5×107cm.s-
1
[113]. Yet, when the measured or calculated parameters are used in the simulator without
any modification, the simulated I-V does not reproduce accurately the experimental, since
parameters like ΦSch, μn and vsat may depend on the implementation of the gate, the
heterojunction, the layer structure and other device parameters. Therefore, it is useful to
consider the measured and reported values only as a guiding advice and to use the
parameters as fitting parameters. Each of the parameters has a distinct impact on the I-V
characteristics.
The I-V characteristics can be split into several distinct sections. For VG < VT there is a
negligible current in the device, for VG > VT, there is a region of linear dependence of
current on voltage and a saturation region. In between the linear and saturation regions
there is a transitional “knee” region. In calibrating a device, it is necessary to match the VT,
the “knee” point in the I-V plane and the slope in the saturation region. The slope in the
linear region is given by VT and the “knee”. VT is defined by σi and Wf. Figure 3.7 shows
the impact of several fitting parameters, μe and vsat in subfigure a) and Rc, σs, Wf in
subfigure b), on the “knee” point in the I-V characteristic. E.g. increasing both the mobility
and saturation velocity increases both the slope of the linear region and the saturation
current but, higher vsat causes the current to saturate at lower VG, while higher μe at higher
VG.
The calibration procedure depends on which parameters we know from the measurements
or literature and with what certainty. In general, we proceed from fitting low to high
current. First, by adjusting Wf, we fix VT, then, by adjusting μe0, we fit the linear regime
and finally, by adjusting vsat, we fit the saturation current. However, the different parts of
the I-V characteristics are not independent and most parameters have an impact on more
3.4 Calibration of the Simulator 55
than one part. Therefore, this procedure is not straightforward and has to be iterated until
reproducing the experimental I-V characteristics.
a) b)
Figure 3.7: Impact of various fitting parameters on I D-VG with focus on the transition from linear regime to
saturation. The parameter that is changed between two simulations is enclosed in a box. The units of σs are %
of what is expected from (2.16) not taking the surface states into account, as discussed in 2.4.2.1. a) The low
field mobility (green line) and the saturation velocity (blue line) are changed with respect to the simulation
represented by the red line. b) The change of the work-function translates to a shift along the VG axis. The
impact of the surface charge and the contact resistance is illustrated as well.
Taking into account Figure 3.7, we can view the impact of the fitting parameters as vectors
in the I-V plane, shifting the transition region. However, the shift depends on VD and the
values of other parameters. The combined impact of the parameters is non-linear but, it is
possible to take advantage of the view of the fitting parameters as operators that shift the
transition region of I-V characteristics.
Figure 3.8 illustrates the calibration method using this transition shift model for I-V curves
shown in Figure 3.7 a). For simplicity, here we assume that only three parameters are
calibrated and keep the rest of the parameters constant.
1. Identify the linear and saturation part of the experimental ID-VG characteristics and,
by the means of linear regression, find the straight lines that approximate both
parts. Calculate the point where the lines cross and label this point in the I-V space
as M.
3. Calculate the crossing-point for the initial simulation as in step 1 and label this
point as S0. Label a vector v0 = M – S0, which is the shift of the transition region
necessary to calibrate the measured data.
4. For each of the fitting parameters, perform a simulation, where this parameter is
changed with respect to the initial simulation and keep other parameters constant.
Label the change of the parameter Δpi, where i is the parameter iterator.
5. For each of the simulations performed in step 4, calculate the crossing-point and
label this point as Si. Calculate vectors vi = Si – S0.
6. Calculate coefficients αi, so that Σ αivi = v0. Since we are fitting three parameters
and the ID-VG space is two dimensional, there is freedom in setting one of the
parameters. Therefore, we may set several values of one parameter and calculate
other parameters for all set values.
7. Perform simulations with the values of parameters changed by αiΔpi for all sets of α
calculated in step 6.
It is advisable to find more than one set of parameters, since, using this method, one only
finds parameters that reproduce the transition point, not the saturation slope. Having more
than one calibration, we can choose the one that is closest to the experimental I-V
characteristics. Moreover, the calibration is done only for one value of VD and may be off
at other VD. To avoid this problem, it is best to search in two ID-VG planes, at two VD
simultaneously, since this reduces the freedom of calculating the parameters in step 6 and
thus leads to increased precision or it is possible to search for more than three parameters
at once.
3.4 Calibration of the Simulator 57
Figure 3.8: Calibration procedure employing the view of the fitting parameters as operators shifting the
transition region. The ID-VG simulations (dashed lines) are taken from Figure 3.7 a). The linear and
saturation regions are identified (solid thick lines) and straight lines (solid thin lines) that approximate the
selected I-V regions are found. A crossing-point (empty circles) is calculated for each pair of the straight
lines. These points are then used to calculate by how much we must change the parameters in the next step
to shift the transition point to that of the experimental curve.
Figure 3.9 shows the cross section of the devices simulated in this thesis. Let us label the
two devices simulated in this work as Device A and Device B. The structure of the devices
is 23nm thick AlxGa1-xN barrier, with x ≈ 28%, on 1.9μm thick GaN layer, which is on top
of a SiC substrate. The source-drain separation is 4μm, the source-gate separation is 1μm,
and the Ni/Au gate length is 0.25μm. The GaN layer is Fe-doped, which acts as a deep
level acceptor with the energy level ≈1eV below CB, for punch-through suppression [114].
Fe concentration is 1×1016cm-3 at the AlGaN/GaN interface, increasing to 1×1018cm-3 at
1μm deep and then constant to SiC substrate. As mentioned in 3.1.2, high concentration
donor doping is used in a small region around the ohmic contacts (source and drain) of the
device in the simulator, to emulate the metal spikes. The width of Device A and B is
2×50μm and 4×125μm, respectively. The devices were produced by the same process, with
the main difference that the wafer, on which Device B was fabricated, shows more
sensitivity to stress and larger DC-RF dispersion, probably because of lost control during
processing [115]. It was taken to our advantage and, in this work, in Chapters 5 and 6, the
stress sensitive Device B was used in investigation of the current collapse and device
3.4 Calibration of the Simulator 58
Figure 3.9: Schematic illustration of the two HEMT devices simulated in this work (not to scale). The
structure of the devices is 23 nm thick Al0.28Ga0.72N barrier on 1.9μm thick GaN, which is on top of a SiC
substrate (not shown here). The source-drain separation is 4μm, the source-gate separation is 1μm, and the
Ni/Au gate length is 0.25μm. The electrodes, i.e. the source, the gate and the drain, are represented by thick
black lines. The light orange shading in the GaN layer represents Fe doping. The abrupt change in the total
polarization, spontaneous (2.2.1) and piezoelectric (2.3.2), represented by the red arrows, gives rise to the
bound charge (2.3.3), represented by violet circles with + and – signs, which induces the 2DEG (2.3.4),
represented by the green dashed line, which makes the operation of a HEMT possible (2.4.1.1).
Certain regions of the I-V characteristics are affected mainly by some parameters and the
fitting parameters usually affect certain regions more than others. This fact is employed in
the calibration of a device, exemplified here on Device A, in Figure 3.10. This figure
illustrates the extraction of three parameters from certain regions of the I-V characteristics.
It is the Schottky barrier height ϕSch from the threshold region of ID-VG, the contact
resistance Rc from the slope of ID-VD close to VD = 0V and the low-field mobility μn0 from
the linear region of ID-VG at low VD. The calibration of μn0 is not straightforward, since
fitting the saturation velocity vsat and the surface charge σs disrupts the already calibrated
mobility. Therefore, calibration of these parameters is an iterative process. The final
3.4 Calibration of the Simulator 59
calibration, reported in Figure 3.12 for a wide range of VG and VD, yielded higher μn0 and
vsat than reported in Figure 3.10 c), which is due to the higher calibrated value of σs. The
calibrated parameters are listed in Table 3.2.
a) b) c)
Figure 3.10: Extraction of some of the parameters in the calibration of Device A. a) Matching the VT in the
simulations with the experimental data determines the Schottky barrier height ϕSch. b) ( ⁄ ) is
affected by the contact resistance Rc. c) Fitting the linear region of ID-VG characteristics yields low-field
mobility μn0. However, to fit the ON-current too, it is necessary to change the saturation velocity, which
disrupts the already fitted part of the curve. The same is true for the surface charge. Therefore, it is necessary
to calibrate the device in an iterative process. I-V characteristics for wide range of gate and drain voltages of
the calibrated Device A is shown in Figure 3.12 with the values reported in Table 3.2.
Device B was calibrated using the procedure described section 3.4.2 and the calibration of
this device is documented in Figure 3.11 with the parameters used in each of the
simulations listed in Table 3.1. The initial set of simulations is shown in subfigure a). First,
a simulation is performed with a set of parameters, labelled orig, and a simulation for each
parameter in which that parameter is changed with respect to orig, while other parameters
are kept constant. In this case, the fitting parameters are the mobility, the surface charge
and the saturation velocity, while the contact resistance (Rc = 750Ω.μm) and the Schottky
barrier height (ϕSch = 0.95eV) were held constant. Then, a crossing point between the linear
and saturation region of ID-VG characteristics is extracted for each of the simulations, as
well as for the experimental data, illustrated in subfigure b). It is possible to calculate the
modification of each parameter necessary to reach the crossing point of the experimental
data. For three parameters, there is an infinite number of combinations but, for a given
modification of one parameter, the two other are determined. The impact of increasing the
mobility on the crossing point is almost the same as the impact of decreasing the surface
charge, but other parts of the I-V characteristics are affected differently by these two
parameters.
3.4 Calibration of the Simulator 60
a) c) e)
b) d) f)
Figure 3.11: Calibration process of Device B, follows the procedure described in section 3.4.2. The units of
the parameters are cm2.V-1.s-1 for mobility, % of the surface charge according to equation (2.15), i.e. without
the effect of the surface traps, equal to -3.4×1013cm-2 and ×106cm.s-1 for the saturation velocity. a) initial
simulation orig with estimated fitting parameters and simulations with one parameter changed each, mob
(mobility), surf (surface charge σs) and v_sat (saturation velocity). b) the corresponding crossing points. c)
and d) show the ID-VG characteristics at VD = 3V and VD = 10V, respectively, calculated from the points in
b). As is obvious, the curves differ negligibly at V D = 3V, which validates the procedure. Using this method,
we can generate a subspace of parameters that give good agreement at a particular voltage and then select a
specific combination based on curves at a different voltage. e) and f) show I D-VG characteristics after several
iterations, simul 31 was selected as the best fit. I-V characteristics using these parameters for a wide range of
VD and VG are reported in Figure 3.13.
The saturation velocity will change very little, since the direction of its impact is almost
perpendicular to the desired shift of the crossing point. The subfigures c) and d) show
simulation results using parameters generated in this way. Since ID-VG curves at VD = 3V
were used as a basis for the calculations, the new curves at this drain voltage, shown in the
subfigure c), virtually overlap. This is not the case at VD = 10V, shown in the subfigure d).
At the higher voltage, the characteristics follow the same path up to a certain VG and then
deviate. The subfigures e) and f) show the simulated characteristics after a few iterations.
The first set of simulations (subfigure d) offer a better calibration for VG ≤ -2V, but
overestimate the current between 1V and 1V and thereafter saturate abruptly, what is not
seen in the experimental data. The final calibration (subfigure f) underestimates the current
for VG ≤ -2V but provide a better calibration for VG close to 0V, when the device is ON and
the saturation appears less abruptly. I-V characteristics using this set of parameters, listed
3.4 Calibration of the Simulator 61
in Table 3.2, for a wide range of VD and VG are compared to the experimental data and
reported in Figure 3.13.
Table 3.1: Fitting parameters used in the simulations of Device B, reported in Figure 3.11. The best fit was
achieved in simul 31, printed in bold font.
a) b)
Figure 3.12: Calibration of Device A. VT ≈ -5V. The parameters used in the calibration are as follows. The
low-field mobility μn0 = 1350cm2.V-1.s-1, the saturation velocity vsat = 1.04×107cm.s-1, the contact resistance
Rc = 350Ω.μm, the Schottky barrier height ΦSch = 1.2eV, the charge at the interface σi = 1.28×1013cm-2 and
the charge at the surface σs = -4.76×1012cm-2.
3.4 Calibration of the Simulator 62
a) b)
Figure 3.13: Calibration of Device B. VT ≈ -5.2V. The
parameters used in the calibration are as follows. The
low-field mobility μn0 = 1450cm2.V-1.s-1, the saturation
velocity vsat = 9.53×106cm.s-1, the contact resistance
Rc = 750Ω.μm, the Schottky barrier height
ΦSch = 0.95eV, the charge at the interface
13 -2
σi = 1.28×10 cm and the charge at the surface σs = -
3.65×1012cm-2
c)
Figure 3.12 and Figure 3.13 show the calibration of Device A and B, respectively. The
accuracy of these results is discussed in the following subsection. Here we focus on the
values of the fitted parameters, summarized in Table 3.2, and compare them with values
reported in literature.
The dispersion of reported values of μn0 is large, e.g. 1460cm2.V-1.s-1 [116] and
2000cm2.V-1.s-1 [112] have been demonstrated in AlGaN/GaN HEMTs by Hall
measurements and 1070cm2.V-1.s-1 [23] and 1700cm2.V-1.s-1 [56] were used in device
simulations, reported to be in agreement with Hall measurements too. The values of the
low-field mobility that we have obtained, i.e. 1350cm2.V-1.s-1 and 1450cm2.V-1.s-1 for
Device A and Device B, respectively, are well within this wide range of reported values.
The situation with vsat is similar, by Monte Carlo calculations it was predicted to reach
2.5×107cm.s-1 [113], while measurements continuously show significantly lower values,
1.1×107cm.s-1 [117] and 1.32×107cm.s-1 [118] and the range used in simulations by various
authors is large too, from 1×107cm.s-1 [23] to 2.3×107cm.s-1 [56]. The values that we have
obtained from the calibration are on the lower end of the range of the reported values,
1.04×107cm.s-1 and 0.953×107cm.s-1 for Device A and Device B, respectively. The
dielectric constant for AlGaN was extracted from the calibration of the dependence of the
3.4 Calibration of the Simulator 63
an impact on some of the other calibrated parameters. Higher Rc would result in higher μn0
and vsat.
Table 3.2: The values of the fitting parameters in the calibrated devices and in the literature.
When we calibrate a device, we do not check the ID-VG characteristics for all measured VD
and ID-VD characteristics for all measured VG every time a parameter is changed, since this
would be overly time consuming. However, it is necessary to perform the calibration using
at least two I-V curves, otherwise it is possible to achieve an excellent fit for a particular
VD and reproduce the ID-VG very accurately, yet at a different VD, the simulated ID-VG may
be very far from the measured characteristics. Usually, we compare the simulated and
measured data for ID-VG at low and high VD (e.g. at 3V and 10V) or at one ID-VG (e.g. at
VD = 3V) and one ID-VD (e.g. at VG = 0V). In this way we ensure a reasonably good
calibration of the drain current for a wide range of both the gate and the drain voltages.
3.4 Calibration of the Simulator 65
Nevertheless, after finding the fitting parameters of a calibration in this way, it is still
necessary to estimate the calibration error for the range of VG and VD of interest.
a) b)
c) d)
Figure 3.14: ID-V characteristics of Device A. a) The measured dependence of the drain current ID on the
drain and gate and voltage, VD and VG. (Figure 3.12 (a) and (b) combined). Calibration error of the
simulated ID, expressed in % above/below the measured values, (b) as a function of both VD and VG and
separately, as a function of (c) VG and (d) VD.
As a measure of the accuracy of calibration, we introduce the calibration error term, which
is calculated as ( )⁄ , where the indices s and m represent the
simulated and measured current, respectively. The measured ID (a) and the calibration
error, expressed in the figures in %, (b) as a function of VD and VG are shown in
Figure 3.14 for Device A and in Figure 3.15 for Device B. The subfigures (c) and (d) show
ΔID as a function of VG and VD, respectively. The different views of ΔID allow to indicate
the regions of the VD-VG plane that are well calibrated. For the calibration of Device A, the
error is less than 3% for VD ≥ 5V and VG = (-3 … 0) V As it is clear in Figure 3.12, the
transition from the linear region is smoother in the experimental data than in the
3.4 Calibration of the Simulator 66
simulations, which leads to higher error in that region. This is represented by the ridge of
high values in Figure 3.14 b) that starts for low VG and VD and shifts to higher voltages at
both electrodes simultaneously. For the calibration of Device B, the error is less than 3%
for VD ≥ 2V and VG = (-1 … 2) V with the exception of six bias points. By comparing the
subfigures (c) and (d) of Figure 3.14 and Figure 3.15 we can infer that, since the error
dispersion is lower in the Device B, that the calibration of this device is more accurate.
a) b)
c) d)
Figure 3.15: ID-V characteristics of Device B. a) The measured dependence of the drain current ID on the
drain and gate voltage, VD and VG. (Figure 3.13 (a) and (b) combined). Calibration error of the simulated ID,
expressed in % above/below the measured values, (b) as a function of both VD and VG and separately, as a
function of (c) VG and (d) VD.
The calibration is accurate if ΔID is close to zero; large negative or positive values are
undesirable. To compare the accuracy of the calibration of the two devices, we show the
average of absolute values of ΔID for range of VD and VG values for both devices in
Figure 3.16. For Device A, the error at VG = -5V was large (almost 25%) and is out of
range of the subfigure a). However, considering that VT ≈ -5V in Device A and ID
3.5 Summary 67
(VT,A) ≈ 0, a large percentage error still means a negligible error in the absolute values.
Therefore, this value is excluded from the calculation in subfigure b). In Device A, the
average error for all values of VD is below 6%. In Device B, the error minimum is shifted
to slightly higher values of VG. For VG between -3 and 0V in Device A and between -1 and
2V in Device B, the average error is below 2.5% and 3.5%, respectively.
a) b)
Figure 3.16: The calibration error as a dependence of a) VG and b) VD, averaged over a range of simulated VD
(1 – 10 V) and VG (-4 – 1 V) points, respectively. The error for VG = -5V (almost 25% for Device A), since it
is close to VT, is excluded from the calculation in subfigure b). The average error for all values of VD is below
6%. For VG between -3 and 0V in Device A and between -1 and 2V in Device B, the average error is below
2.5% and 3.5%, respectively.
3.5 Summary
This chapter provided an overview of the Sentaurus simulation platform from Synopsys
with an introduction to its tools that were most important for this work, the Sentaurus
Workbench, Sentaurus Structure Editor and Sentaurus Device. Further, the most important
physical concepts were briefly introduced, such as mobility, drift and diffusion currents, as
well as the equations that are solved by Sentaurus Device tool in the drift-diffusion model,
such as the Poisson‟s equation, current continuity equations and the expression for the
drift-diffusion current. An important ingredient of this thesis was to automate the
simulation process to perform large number of simulations with predetermined sets of
values for various fitting parameters, to perform simulations in a loop, in which the
following run depended on the results from the previous run, to perform calculations
3.5 Summary 68
outside the scope of the Sentaurus simulation platform and couple those calculations to the
simulations and, finally, to evaluate the results from all performed simulations. To
accomplish these tasks, various procedures and scripts were developed and described
concisely in section 3.3. The first indispensable step in investigating any as yet not fully
understood aspect of a phenomenon with the help of physical simulations, is to calibrate
the simulator against the experimental data in well-known conditions, thus validating the
model, the implementation, i.e. the simulation tool, the applicability of the model and the
used parameters. In the case of transistors, this is usually accomplished by the calibration
of I-V characteristics of the device, discussed thoroughly in section 3.4. First, we introduce
the fitting parameters used in the calibration with values reported in the literature, describe
the method used in the process of calibration, where we discuss the impact of several of the
parameters on different regions of the ID-V dependence, and finally, we give an account of
the calibration of the two devices used throughout this thesis. Then, we give a detailed
description of the devices, extraction of some of the parameters from different parts of the
I-V characteristics, the calibrated I-V for wide range of applied drain and gate voltages, we
discuss the calibrated parameters and compare their calibrated values to values reported in
literature. Finally, we comment on the accuracy of the calibration and estimate the
calibration error, which is below 3% for a wide range of bias conditions, specified in
3.4.3.3, and does not exceed 10%, except for VG close to VT, where ID is very low. The
calibrated devices are used in the rest of the thesis. As mentioned in 3.4.3.1, Device A was
used in the investigation of the polarization induced bound charge and its dependence on
VG in Chapter 4, and Device B was used in the investigation of current collapse and device
degradation phenomena in Chapter 5 and Poole-Frenkel conduction mechanism probably
responsible for DC-RF dispersion in Chapter 6.
The simulations in this thesis were performed on the computing cluster that is used by the
Device Modelling Group at the University of Glasgow. The cluster is a 1232 core cluster
and contains 64 chips (256 cores, 4 cores per chip) of E5462 2.8GHz Xeon with 8GB per
node (1G per core) and 122 chips (976 cores, 8 cores per chip) of E5530 2.4Ghz Xeon with
24GB per node (3GB per core).
The CPU time of a device simulation depends on the size of the mesh, on the range of
voltages and on the convergence of a given problem. The simulation time of single I-V
3.5 Summary 69
simulations (e.g. ID-VG at a particular VD) performed in this thesis typically varied between
ten minutes and one hour. However, to fully investigate a phenomenon or achieve a goal,
such as device calibration, it is necessary to perform number of simulations. In the case of
the device calibration, as reported in Table 3.1, it took 32 iterations plus four initial
simulations, all at two different drain voltages, resulting in total 72 device simulations, i.e.
approximately 36 hours of CPU time. In Chapter 5, in the search for the surface charge
density distribution of a „virtual gate‟ that reproduces the current collapse phenomenon,
two iterations of all 36x4 parameter combinations, reported in Table 5.1, were performed,
resulting in 5832 simulations, which was approximately 3000 CPU hours. And finally, in
Chapter 6, in the investigation of Poole-Frenkel transport mechanism that leads to the
„virtual gate‟ responsible for the current collapse, 250 iterations were performed at high
drain voltage (resulting in longer simulation time) for a single set of investigated
parameters. In total, 76 different combinations of four model parameters were simulated.
This results in 19‟000 simulations, which translates to minimum of 104 CPU hours. The
large cluster was indispensable in achieving the results presented in this thesis.
4 Polarization Induced Bound Charge
4.1 Introduction
Using self-consistent numerical simulation, we have studied the impact of the field-
induced polarization on the characteristics of AlGaN/GaN HEMTs. In the simulations, the
strain induced by external electric fields is added self-consistently to the built-in strain due
to lattice mismatch. It has been suggested that this additional strain can play a role in the
device degradation and failure [29]. The study is carried out using commercial TCAD tool
carefully calibrated against the measured characteristics of 0.25 m physical gate length
AlGaN/GaN transistors. A coupled model for piezoelectric materials, including the impact
of the field, is used to determine the strain and thus the polarization in the device. The
spatial charge distribution is derived from the gradient of the polarization and is fed back
self-consistently to the simulator.
The diagram in Figure 2.3 shows the relationships between physical variables in a
piezoelectric material. Until now, we have only considered the direct piezoelectric effect,
i.e., the electric polarization field induced by the stress applied to a crystal, which is
represented by the left side of the diagram and described by the equation (2.4). By doing
so, we have neglected the impact of the electric field on the stress, strain, and polarization
of the crystal. However, during the operation of a HEMT, a large electric field is induced
4.2 Converse Piezoelectric Effect 71
in the device, especially near the contacts. This field induces additional stress in the device.
To capture the stress, strain, and polarization distribution in the device, the impact of the
electric field has to be included. The right half of Figure 2.3 demonstrates the converse
piezoelectric effect, i.e., the deformation of a crystal induced by the electric field. To take
the converse piezoelectric effect into account, the simple relationship between stress σ and
strain ε, given by (2.2), has to be extended to include the contribution of the electric field.
The fully coupled equation [120] for piezoelectric materials can be read from the diagram
in the Figure 2.3 as
or i Sij j d ki Ek (i, j = 1 .. 6, k = 1 .. 3)
where E stands for components of the electric field vector and all the indices are given by
the Voigt notation, as defined in Table 2.2.
Figure 4.1: Schematic representation of the converse piezoelectric effect, based on the set of equations
(4.1). The crystal direction is shown on the left. The wurtzite structure is asymmetrical in the z direction, so
the z-component of the electric field, Ez, will have opposite effect in the positive and negative directions.
The two rows represent the opposite directions of Ez. Other components of E induce only shear strains. In
all pictures, the thick grey lined square represents the crystal without the effect of E; the thin black lined
square represents the deformation due to the E, as well as external stresses needed to maintain the desired
shape. The three columns represent three different boundary conditions. The first column is in the absence
of any mechanical external forces, σi = 0, for all i. In this case, the crystal simply expands in one direction
and contracts in the other. The second column is for a completely fixed structure, without the possibility to
change the shape. In this case, E will produce forces on surrounding material. The arrows are in the
direction of external forces that need to be applied on the structure to prevent it from deforming. Finally,
the clamped model, given by the equations (4.3) and (4.4), is shown in the third column. The dashed line
represents a freestanding structure, the grey line represents a strained structure, e.g. a thin AlGaN barrier
layer grown on GaN, strained to match the underlying layer, and the black line represents the structure
under the impact of E. The stress shown in this column is a change in the stress already present due to the
layer being already strained.
C13 e
(4.3) 3 2 1 33 E3
C33 C33
Substituting the vertical strain in the expression for the lateral stress we obtain
C132 C
(4.4) 1 C11 C12 2 1 13 e33 e31 E3
C33 C33
In the absence of the electric field, the equations (4.3) and (4.4) are reduced to (2.8) and
(2.10), respectively. Figure 4.1 shows the essence of the converse piezoelectric effect, with
different boundary conditions. The rightmost column shows the clamped model
approximation. A detailed explanation is given in the caption to the figure.
To give an idea about the magnitude of strains and stresses in devices under the impact of
the electric field, we plot them in Figure 4.2 as a function of E for various Al fractions x.
To demonstrate the fact that the electric field always increases the strain state of the
crystal, the strain energy per unit volume, which is calculated as [46, p. 136-137]
1
(4.5) W Cij i j ,
2
is plotted as well. The part of the equation (2.4) that holds, even after taking the converse
piezoelectric effect into account, is the relation between the piezoelectric polarization Ppz
and strain ε. Using the Voigt notation, the relation can be rewritten as [121]
(4.6) Pi pz eij j
4.2 Converse Piezoelectric Effect 74
a)
c)
b)
Figure 4.2: Subfigures a) and b) show vertical strain ε3 (4.3) and lateral stress σ1 (4.4) respectively, in
accordance with the clamped model, as a function of the z-component of the electric field E3, for various Al
fractions x of AlGaN, assumed to grow on a thick relaxed GaN layer. From the subfigure a) it may seem that
an electric field parallel with the z direction (E3 > 0) reduces the strain. Nevertheless, we must remember that,
the layer grows with a built-in lateral strain ε1; the vertical strain ε3 is a result of the assumption of no force
applied in the z direction. Since the electric field in any direction exerts an additional force, any variation in
strain will only increase the total strain of the crystal. This is demonstrated by the subfigure c) in which strain
energy per unit volume versus the vertical strain or electric field is plotted. A non-zero electric field can
increase or decrease the vertical strain, but it always increases the strained state of the crystal.
Considering the expression for the vertical strain ε3 given by (4.3), the z component of the
piezoelectric polarization can be expressed as
pz C13 e332
(4.7) P3 2 e31 e33 1 E3
C33 C33
Again, in the absence of the electric field (or by neglecting its contribution), we arrive at
the previously derived equation (2.9). We will label the part of the piezoelectric
polarization that is independent of the electric field as P0pz . It can be the first term in the
4.2 Converse Piezoelectric Effect 75
formula (4.7), which is identical to equation (2.9), in the model of linear dependence of
piezoelectric polarization on strain, or the combination of the equations (2.11) and (2.12)
for the non-linear model, as discussed before.
The clamped model is an approximation and as such has its limitations, which are a
consequence of the simplifying assumptions behind the model. The strain / stress in a
particular location depend only on built-in strain and the z component of the electric field
in that location. Apart from ignoring the impact of other components of the electric field
and the resulting shear strain / stress, the model neglects mechanical forces exerted by the
surrounding material. The non-equilibrium situation produced by this model in a region
with Ez rapidly changing in the x direction is shown in Figure 4.3.
Figure 4.3: The deficiency of the clamped model comes from its simplifying assumptions is shown by
considering two adjacent elements of a piezoelectric material. Dashed lines represent the situation before
applying the electric field. The electric field in the “element 2” is greater than that in the “element 1”. From
(4.3) and (4.4) it follows that, the lateral stress and vertical strain in those two elements will be different.
Two obvious problems arise from this result. One is non-equilibrium in stress; the “element 2” will press
on the “element 1” with larger force than the other way around. The second is that a point on the top
boundary of the two elements (full circle) will split in two under the influence of the electric field. In a
solid matter, this is not possible. As a consequence, even by neglecting (or in the absence of) the x and y
components of the electric field, there will still be shear strains and stresses in the device.
During the device operation under bias, there is a spatially varying electric field. This has
two important consequences. Firstly, according to the equation (4.3), there will be spatially
varying strain field in the device, which can lead to a change in device properties [122]. In
4.2 Converse Piezoelectric Effect 76
addition, a long-time (several hours) operation under high electric and hence strain field
may result in defect formation and therefore in the device degradation [29]. Secondly,
according to the equation (4.7), there will be varying polarization field in the device. If we
link this result with the fact that, the divergence of the electric polarization is associated
with the bound charge (2.14), we realize that, apart from the discontinuity at the interface,
as given by (2.15), there will be bound charge also in the bulk of the device. After
substituting the expression for the piezoelectric polarization Ppz (4.7) in the differential
form of (2.14), we obtain the expression for the bound space charge as
e332 dE3
(4.8) b
C33 dz
Figure 4.4: The impact of the converse piezoelectric effect on the bound sheet charge at the heterojunction
interface. The situation on the left is without taking the effect into account. The spontaneous polarization is
not shown, because it is not affected by the electric field or the additional strain. The gray square represents
the unstrained AlGaN. AlGaN grows with strain on the relaxed GaN, which results in piezoelectric
polarization in AlGaN, and according to (2.15), the difference in the polarizations leads to formation of the
interface charge. The charge induces the electric field E and, via the converse piezoelectric effect, produces
additional strain in both AlGaN and GaN. This modifies the polarization in both layers and hence alters the
bound charge, as derived in (4.9). The bias applied at the electrodes, especially the gate, modifies the
electric field and therefore the bound interface charge.
Moreover, the bound sheet charge σb at the interface will be modified as well. Substituting
(4.7) into the general formula for σb (2.16), we obtain
e332 x e 2 0
(4.9) b P sp 0 P sp x P0pz x E3 AlGaN 33 E3 GaN
C33 x C33 0
4.3 Simulation Methodology 77
The first three terms are those that are present in the formula even without taking the
converse piezoelectric effect into account. The last two terms arise due to the converse
piezoelectric effect. The E3 in the formula is the electric field in the two materials, right
next to the interface. The impact of the electric field via the converse piezoelectric effect is
shown in Figure 4.4.
The first step in the self-consistent cycle is to perform simulations without any additional
electric field-induced charge. This was used to calibrate the TCAD simulator in respect to
the measured ID-VG characteristics of a real HEMT achieving the excellent agreement as
illustrated in Figure 3.12. Initially, we consider only the built-in polarization, i.e., the
spontaneous polarization and the piezoelectric polarization originating from the strain of
the Al0.28Ga0.72N barrier only, which results in the interface charges and induces carriers in
the channel. At this stage, the spatial distribution of the electric field and the corresponding
field induced polarization and space charge in the two materials are excluded from the
solution of the Poisson equation.
Based on this initial simulation, the electric field is evaluated at all locations in the device.
Then, using the electromechanical coupling (4.3), this electric field is linked to the strain.
Subsequently, this strain is linked to polarization (4.6) and, finally, the polarization to the
charge distribution in the device, while using equation (4.8) inside the device and equation
(4.9) to modify the charge at the interface of the device. Charge distribution is then fed
4.3 Simulation Methodology 78
back into the simulator and the simulations are performed again. This procedure,
schematically shown in Figure 4.5, is repeated until convergence is achieved making the
whole simulation process self-consistent.
Figure 4.5: Simulation flow to show the impact of the converse piezoelectric effect on the simulated I-V
characteristics and electric field distribution in the device.
4.4 Results 79
4.4 Results
Figure 4.6: Distribution of the electric field in the device at drain voltage V DS = 3V and gate voltages
VGS = -6V and VGS = 0V, just under the threshold.
Figure 4.7: The electric field E distribution near the gate in the GaN layer, 0.1 nm under the interface, in
the linear regime, VD = 3V, and saturation, VD = 20V. The position of the gate is indicated by the thick
solid line at the bottom of each graph and vertical thin dashed lines on the edges of the gate.
4.4 Results 80
The first stage in the simulation loop is to simulate the device (especially the I-V
characteristics and the electric field) considering the built-in strain and the resulting direct
piezoelectric effect only, i.e., for i = 0 in Figure 4.5. Figure 4.6 shows the initial
distribution of the electric field in the transistor, at two different gate voltages (VG = 0V, at
the on-current, and VG = -6V, which is just below the threshold) at drain voltage VD = 3V.
Figure 4.7 shows the x and z components of the electric field along the channel, 0.1 nm
under the AlGaN/GaN interface, for three different gate voltages: V G = 0V (ON-current),
VG = -4V (linear regime), and VG = -6V (OFF-current). Ex is decoupled from the clamped
model since it induces shear strains only and they are neglected in the model. However, Ez
alters the bound charge distribution in the device, which has an impact on the electrostatic
potential distribution and therefore on Ex. So, as a result, Ex will change as well. According
to equation (4.3), the vertical electric field induces strain in the device, in addition to the
built-in strain that is already present due to lattice mismatch between the layers. This
additional strain is displayed in Figure 4.8 at low drain voltage, VD = 3V, VG = -6V, and in
Figure 4.9 and Figure 4.10 at high drain voltage, VD = 20V, in off- (VG = -6V) and on-state
(VG = 0V) of the device, respectively. The strain, in turn, induces spatially varying
piezoelectric polarization, which gives rise to bound charge, given by equation (2.14) in
general or (4.8) and (4.9) in the case of wurtzites using the clamped model, in both off- and
on-state conditions. Figure 4.11 and Figure 4.12 show the distribution of the bound charge
near the gate at low (3V) and high (20V) drain voltage, respectively. Feeding this charge
distribution back into the device, the simulation of the ID-VG characteristics was performed
until convergence, as indicated in Figure 4.5. The results of the self-consistent simulations
are analysed in the following subsections.
4.4 Results 81
Figure 4.8: The distribution of change of the vertical strain in the region close to the gate in the off-state
and at a low drain voltage, VD = 3V, VG = -6V.
Figure 4.9: The distribution of change of the vertical strain in the region close to the gate in the off-state
and at a high drain voltage, VD = 20V, VG = -6V.
Figure 4.10: The distribution of change of the vertical strain in the region close to the gate in the on-state
and at a high drain voltage, VD = 20V, VG = 0V.
4.4 Results 82
Figure 4.11: Distribution of the polarization induced charge in the device, close to the gate, at V DS = 3V.
Figure 4.12: Distribution of the polarization induced charge in the device, close to the gate, at V DS = 20V.
4.4 Results 83
Figure 4.13: Simulation of ID-VG characteristics (top) in the linear region of the device, at V D = 3V. The
uncoupled simulation is without the contribution of the converse piezoelectric effect. The simulation
labelled as coupled (space) takes the contribution of the bound space charge into account, and the
simulation labelled as coupled takes also the modification of the bound sheet charge at the interface into
account. The two bottom graphs show the drain current shift in the coupled (space) (left) and coupled
(right) models, with respect to the gate voltage VG. In saturation regime, the contribution of both effects is
comparable, while for the linear regime and at the threshold voltage, the modification of the bound sheet
charge at the interface is dominant.
The converse piezoelectric effect results in (i) induced bound space charge in the device,
given by (4.8), which was studied in [123,124], and in (ii) modified bound sheet charge at
the interface, given by (4.9). In order to determine how do these two effects contribute to
the simulation of the operation of the device, i.e., whether one of them is dominant or they
are comparable, we ran simulations at VD = 3V, i.e. in the linear region of the device.
4.4 Results 84
The simulation was run in three conditions (i) uncoupled: neglecting the converse
piezoelectric effect, which is the standard method in the most simulations of III-N devices
(ii) coupled (space): taking only the bound space charge into account and finally (iii)
coupled: taking both the bound space charge and modification of the bound sheet charge
into account. The ID-VG characteristics for all three simulations are shown in Figure 4.13.
In the two bottom subfigures, we can see that the impact of both of the coupled conditions
increases with increasing the negative gate voltage, because of the higher electric field. In
the saturation region, at VG ≈ 0V, the spatial bound charge increases the drain current by
approximately 0.5%. When the modification of the sheet charge is added to the simulation,
the overall effect is reduction of the ID by approximately 0.5%. Hence, the effect of the two
types of the bound charge is of similar magnitude, but the sheet charge modification
overrides the effect of the space charge. However, at higher negative gate voltages,
especially close to the threshold voltage, i.e., VG ≈ VT, the sheet charge effect becomes
dominant, and the space charge can be neglected. The main impact of the change in the
bound sheet charge is a threshold voltage shift. The gate voltage, at which the saturation
region is reached, is virtually unchanged. For the rest of this chapter, all the coupled
simulations will take both the bound space and sheet charge into account.
Figure 4.14 shows how the bound sheet charge is changed at the interface due to the
converse piezoelectric effect, as given by (4.9). In order to investigate how this effect
depends on the drain voltage, we performed the self-consistent simulations at high drain
voltage (VD = 20V) as well. Figure 4.15 shows the impact of the converse piezoelectric
effect on the ID-VG simulated characteristics. The effect is essentially the same as for low
drain voltage (VD = 3V, already reported in Figure 4.13). To make an insight, the drain
current change at both simulated drain voltages is included in the figure.
4.4 Results 85
Figure 4.14: Modification of the bound sheet charge σb at the interface. Away from the gate, the electro-
mechanical coupling results in a constant reduction (approx. 1.5%) of the charge, irrespective of the voltage
applied. Under and close to the gate, the electric field is strongly modified by the voltage applied at the
electrodes, especially at the gate. Therefore, the bound sheet charge is modified as well.
Figure 4.15: The impact of electro-mechanical coupling on the ID-VG characteristics at a high drain voltage,
VD = 20V. The effect is virtually independent on the drain voltage. To give an insight, the drain current
shift for low drain voltage (V D = 3V, from Figure 4.13) is included in the figure as well. Again, the main
effect is in the threshold voltage shift, and diminishes with increasing the gate voltage.
4.4 Results 86
Figure 4.16: The electric field E distribution along the channel, under or close to the gate, 0.1 nm under the
heterojunction interface. The top subfigures show the electric field change at low drain voltage, the bottom
subfigures at high VG. Ex is shown on the subfigures on the left, Ez on the right. The effect of the
polarization induced bound space charge alone (top: coupled (space)) on the simulation of E is negligible.
At gate voltages below threshold (left), the change of Ex is independent of VD.
The polarization induced charge and its modification due to the electro-mechanical
coupling has an impact on the electric field distribution in the device. To illustrate this
effect, in Figure 4.16 we plot the x and z components of the electric field along the channel,
0.1 nm under the AlGaN/GaN interface at both low (VD = 3V) and high (VD = 20V) drain
voltages. At VD = 3V, the distinction between the two coupled models (induced space
charge vs. space and sheet charge) is made. As expected from the already shown I-V
characteristics, the effect of the induced space charge on the electric field is negligible.
However, at high negative gate voltages, the fully coupled model causes an increase of Ex
at the source edge of the gate, which in effect raises the barrier for electrons. Hence the
threshold voltage shift observed in Figure 4.15.
4.4 Results 87
Figure 4.17: The conduction band (CB) in the channel, 0.1 nm under the heterojunction interface. The top
subfigures show the CB below the threshold voltage of the device, where the potential barrier prevents
electrons from flowing through the channel and hence the channel is closed. For both of the considered drain
voltages, the barrier is increased. In the bottom subfigure, the CB shift due to the converse piezoelectric
effect for various bias conditions is summarized. For voltages below and around the threshold voltage, the
region that has an impact on the device operation is under the gate, since it determines the height of the
potential barrier. For voltages above the threshold, the important region is between the source and gate. With
increasing the gate voltage, the CB shift in this region fades away.
which the channel closes). At high VD, the largest CB shift due to piezoelectric coupling
appears at the drain edge of the gate. But this is already behind the potential barrier, and
the electrons, if there were any, will be accelerated towards the drain anyway. Therefore,
the CB shift in this region plays no role. However, under the gate, the CB shift is similar to
the conditions at low VD. At that point, where the potential barrier along the channel is the
highest, the CB shift at VD = 20V is approximately equal to 0.14 eV while, at VD = 3V, it is
0.13 eV. Hence, the resulting VT shift is virtually the same for all voltages. The shift at the
low drain voltage (VD = 3V) and when the channel is open channel (VG = 0V) was virtually
zero, which resulted in no change in the drain current.
4.5 Summary
The effect of the electric field induced strain on the I-V characteristics at conditions
considered in this chapter is not dramatic. At higher voltages, the strain induced by the
converse piezoelectric effect may lead to defect formation. The main purpose of this
chapter is to illustrate a self-consistent approach to couple the electric field with the space
charge as advancement to the previous approach that does not take the converse effect into
account.
5.1 Introduction
Current collapse and device degradation in GaN HEMTs associated with charge movement
and trapping in the device present a serious reliability problem. Current collapse has been
analysed previously using numerical simulations [97], and was explained in the context of
a „virtual gate‟ [116,34], associated with additional charge trapping at the interface or in
the bulk regions of the transistors. The proposed explanation of the device degradation
mechanism [29] is that it occurs due to creating lattice defects by excessive mechanical
stress caused by converse piezoelectric effect in the regions with high electric field. These
defects act as traps. In this chapter, we aim to find the surface or volume charge
distribution that would reproduce the I-V characteristics during the current collapse and the
I-V characteristics of a degraded device after a high voltage stress test run for 12.5 hours.
In section 5.2 we explain the operation mode (measurement technique), at which the
current collapse phenomenon occurs. Then we report the results of our investigation of this
phenomenon. First, we investigate the impact of the trapped charge on the I-V
characteristics, by uniform slabs of charge, in section 5.3.1. Next, in section 5.4, we obtain
exponential distributions on both sides of the gate on the device surface that provides a
5.2 The Current Collapse and Device Degradation Phenomena 90
reasonable fit to the experimental I-V measurements pulsed from a class AB point. In
subsection 5.4.2 we explain the automated procedure to obtain the fitting.
In section 5.5 we identify areas of the device that undergo an excessive stress due to
converse piezoelectric effect at high voltages and, by introducing trapped charge in those
areas, attempt to reproduce I-V characteristics of a degraded device.
The current collapse and device degradation have similar impact on the I-V characteristics,
both reduce the current flowing through the channel, especially in the knee region of the I-
V plane. The main difference, as described in section 2.5.2, is that the current collapse is
reversible and the degradation is not. The difference in the mechanism is that the current
collapse is caused by trapping of the electrons in existing traps while the device
degradation is caused by creating new defects and hence new traps in the device, on which
charge is permanently trapped.
During a pulsed measurement, the device is pulsed to a measurement voltage (VG and VD)
for time tm. In between the pulses, the device is held at a quiescent bias (at a gate voltage
VG,q, and drain voltage VD,q) for time tq. This is repeated for how many VG-VD
combinations the current is measured. tq is typically several orders of magnitude larger
than tm. The pulsed I-V characteristics of a device depends not only on the actual
measurement voltage V, but on the quiescent bias Vq, too. If the quiescent bias is zero volts
on both the gate and the drain electrodes, VG,q = 0V and VD,q = 0V, it is called the open-
channel condition. At a different quiescent bias, the current during the pulse will change.
This change in the pulsed-current is referred to as the current collapse that causes the DC-
RF dispersion and is illustrated in Figure 5.1. This effect is caused by electron leakage
mainly to the surface and also to the bulk of the device in between the pulses, as shown in
Figure 5.2, at Vq.
5.2 The Current Collapse and Device Degradation Phenomena 91
a) b)
Figure 5.1: c) An illustration of the dependence of the
measured current on the quiescent bias at which the device is
hold before the actual measurement. The real a) ID-VG and b)
ID-VD characteristics for two different quiescent bias voltages.
The term open-channel refers to the quiescent bias of
VG,q = 0V, VD,q = 0V; whereas the measurement that showed
c)
the current collapse to VG,q = -4V, VD,q = 25V. The dots
represent the measurement point of VG = 0V, VD = 5V, but the measured current depends on the quiescent
bias, a bias applied before the measurement. This phenomenon is referred to as DC-RF dispersion or
“knee-walkout” due to its representation in the I-V plane (blue arrow), as mentioned in section 2.5.2. The I-
V characteristics for a degraded device in the open-channel condition is shown in blue lines.
During the stress test, a device is subjected to high electric field which, via the converse
piezoelectric effect, discussed in 4.2 and described by equations (4.1) and (4.4), induces
strong mechanical stress in the region of the device with high E, which in turn leads to
additional strain of the material, given by equation (4.3). According to [29] this leads to
defects in the semiconductors‟ lattice structure and therefore to creation of new traps.
In our calibration, we have used three sets of experimental data, shown in Figure 5.1. I-V
characteristics pulsed from an open-channel condition – this was used to calibrate the
device, and from the quiescent bias of VG,q = -4V, and VD,q = 25V – this was used to
calibrate the pulsed I-V characteristics to find a surface electron distribution that would
reproduce the measured current collapse data. The measurement time was tm = 1s, and the
time between the pulses was tq = 1ms. In Chapter 6, this distribution was used to simulate
the actual leakage mechanism. Finally, I-V measurement pulsed from an open-channel
condition after the device was stressed and therefore degraded by high electric field. The
device was degraded during RF power test at VG = -2.7V, VD = 25V for 12.5 hours, in
5.3 Investigation of the Impact of the Surface Electron Distribution 92
which the RF output power dropped to 2W/mm at 2.8GHz from the initial 4W/mm [115].
In our simulations, we have identified the region where, due to the high electric field, a
high stress is expected. Assuming that new traps will be created in this region, we
attempted to reproduce the post-degradation measurement by putting charge with a
corresponding distribution in this region.
Figure 5.2: The mechanism responsible for the current collapse. Due to the strong electric field (red arrow),
through the means of the Poole-Frenkel conduction (green arrow), the electrons leak to the surface of the
device and electrostatically deplete the channel (blue arrow) and hence cause reduction of the saturation
current. This is the primary Poole-Frenkel mechanism. The secondary mechanism is that the electrons at the
surface create additional electric field which forces the later leaked electrons to transfer to the traps in the
bulk and to the AlGaN/GaN interface. During a stress test, in the region of the high electric field (red arrow),
new dislocations, and therefore traps, are created. This leads to a permanently degraded device and reduced
current in subsequent measurements. Adapted from [125]
When the current collapse/knee walkout or degradation is apparent (Figure 5.1 b), the ID-
VD slope in the knee region decreases, whereby indicating an increase in the differential
access resistance. The saturated current is also reduced. A threshold voltage shift present in
the ID-VG characteristics of the current collapse measurement, unlike that of a degraded
device, as shown in Figure 5.1 a), indicated the presence of charge trapped under the gate
in addition to that on the ungated surface. Figure 5.3 shows three regions of the device,
where the trapped charge was placed to capture the current collapse phenomenon. To
reproduce the experimental current collapse and degradation in the simulation, we studied
the impact of various distributions of trapped charge at the surface of the device.
5.3 Investigation of the Impact of the Surface Electron Distribution 93
Figure 5.3: A scheme of the device with focus on the regions (red lines) with the trapped charge to reproduce
the pulsed measurements of the I-V characteristics from the class AB point.
a) b)
c) d)
Figure 5.4: Schematic representation of charge distributions used in simulations investigating the impact of
uniform slabs. The red letter „G‟ represents the position of the gate, each shade of grey represents different
simulation. The corresponding simulated I-V characteristics are shown (a) in Figure 5.5, (b) in Figure 5.7, (c)
in Figure 5.6 and (d) in Figure 5.8.
All figures in this section showing simulation results of I-V characteristics also show the
measurements pulsed from an open-channel condition and those showing the current
collapse effect. However, the insights in terms of the impact of the trapped charge on the I-
V characteristics gained here are valid also for the calibration of the device degradation.
The position and the concentration of the trapped charge and its impact on the I-V
characteristics have been studied, as reported in [126]. In this section, all reported ID-VD
and ID-VG experimental data and simulations were measured or performed at VG = 0V and
VD = 3V, respectively. Another important figure of merit in the analysis of a device is
transconductance gm, which indicates the amount of control the gate voltage has on the
drain current. It is defined as the ratio of the current change at the output port to the voltage
change at the input port, i.e., ( ⁄ ) . As a function of VG, gm has a hat-shaped
5.3 Investigation of the Impact of the Surface Electron Distribution 94
dependence, shown in Figure 3.13 c), which rises sharply around VT, then reaches a plateau
and falls moderately before ID saturates. In the following we illustrate the impact of the
surface charge upon this parameter.
a) b)
Figure 5.5: (a) ID-VG and (b) ID-VD characteristics
for different values of surface electron density, 1-
3×1013cm-2, placed at the edges of the gate, in a
region extending 50nm towards the other electric
contacts. Increasing the amount of charge leads to
reduction of the saturation current, but the slope of
the linear regime of ID-VD remains unchanged. (c)
The impact of the surface charge on the
c) transconductance of the device, the black line
represents the device with no trapped charge.
Initial simulations used rectangular charge slabs [97], which were placed on the source and
drain sides of the gate and varied in length and charge density. First, we assume
symmetrical charge distribution on both sides of the gate. The dimensions and charge
density of the slabs are schematically illustrated in Figure 5.4. The impact of changing the
surface electron density (1-3×1013cm-2) in a region of a constant size, 50nm from the gate
towards the other two electric contacts, is investigated in Figure 5.5. The impact of the
length (30-70 nm) of the region at a constant surface electron density, 2×1013cm-2, is
investigated in Figure 5.7. In both cases, increasing the total amount of charge reduces the
saturation current, but fails to increase the access resistance and hence change the slope of
the linear region, necessary to fit the experimental ID-VD characteristics.
5.3 Investigation of the Impact of the Surface Electron Distribution 95
Figure 5.6 compares a) ID-VG and b) ID-VD characteristics for charge extending 50 nm and
150 nm away from the gate. Charge close to the gate shifts the threshold voltage, while
charge that extends further away from the gate reduces the ON-current. With increasing the
region where the charge is trapped, the saturation ID increases less with increasing VD, yet
the impact on linear region of ID-VG is lowered.
a) b)
Figure 5.6: The impact of charge in regions of two
different sizes (to 50nm and 150nm away from the
gate) on the I-V characteristics (a, b) is compared.
Several values of the electron sheet density are used
for both regions. (c) Transconductance for the
150nm region of the trapped charge (for 50nm
region it is reported in Figure 5.5 c).
c)
If the region of the trapped charge is relatively small and the charge density is high, as will
be demonstrated in Figure 5.10 c), the decline of gm happens over wider range of VG than
without the trapped charge. The value of VG at which gm in the simulation with and without
the trapped charge cross, is determined mainly by the length of the region, as is clearly
demonstrated in Figure 5.6 c), for 150nm length of the region. This nearly holds for 50nm
as well, shown in Figure 5.5 c), although here the cross-point shifts slightly and VT is
affected significantly with varying the surface charge density. The impact of varying the
length of the region while keeping the surface charge density constant, shown in
Figure 5.7, is similar to varying the charge density, but here the cross-point shifts clearly
and significantly.
5.3 Investigation of the Impact of the Surface Electron Distribution 96
a) b)
Figure 5.7: (a) ID-VD and (b) ID-VG characteristics for
different values of the extent of the electron charge
from the gate, 30-70nm, of constant sheet density,
2×1013cm-2. The impact of extending the region of
the trapped charge and increasing the charge density,
shown in Figure 5.5, is similar, but not equal, as (c)
the transconductance dependence shows.
c)
The impact of charge trapped at the AlGaN/GaN interface under the gate is investigated in
Figure 5.8. This charge partly neutralises the bound charge under the gate at the
AlGaN/GaN interface originating from the difference in polarization between the two
materials, as explained in subsection 2.3.3. The impact of this charge on the conduction
band is illustrated in Figure 2.6. This charge hence directly depletes the 2DEG under the
gate and shifts the threshold voltage, VT.
Figure 5.8: The sole effect of charge trapped at the AlGaN/GaN interface under the gate is the threshold
voltage, VT, shift. The slope and the ON-current remain unchanged. Apart from the varying charge under the
gate, there was sheet charge density of 6.5×1012cm-2 extending to 150 nm on both sides of the gate, reported
in Figure 5.6.
5.3 Investigation of the Impact of the Surface Electron Distribution 97
Further on, we investigate the impact of the charge on the drain and source sides of the
gate on ID-VG and ID-VD characteristics, for two regions of trapped charge of different
extent. In Figure 5.9 we have used much larger region and lower charge density of the
trapped charge than previously, 1m on both sides of the gate, and in Figure 5.10 we have
used a previously reported region of 50nm and the sheet charge density 2×1013cm-2 (green
line in Figure 5.5 and Figure 5.7). The 1m region of the trapped charge, shown in
Figure 5.9, was enough to change the access resistance, as seen from the changed slope of
ID-VD characteristics. From the ID-VG graph it becomes clear that both areas have the same
effect on the ON-current, while only the charge on the source side changes the slope in the
linear regime. Therefore, this suggests that, to achieve the change in the slope seen in the
experimental data, we need to include the trapped surface charge at the source side. For
drain voltages of up to VD = 4V, the charge in both areas has the same effect on the linear
regime. For higher drain voltages, the current is limited by the charge on the source side. In
the ID-VD characteristics, the charge at the source side saturates current at lower VD,
whereas the experimental data show continuous increase in the current as in the case when
charge on the drain side is included. From this fact, we conclude that, to reproduce the
experimental slow increase in the saturation current, the trapped charge needs to be mainly
on the drain side of the gate.
a) b)
Figure 5.9: Asymmetric charge distribution around
the gate. The impact of the trapped charge at the
source and drain sides of the gate is investigated
separately, for low charge density in a large region.
(a) ID-VG and (b) ID-VD characteristics and (c)
transconductance are compared.
c)
5.3 Investigation of the Impact of the Surface Electron Distribution 98
The impact of charge trapped on either side of the gate on the gm-VG dependence is most
clearly demonstrated using moderate surface charge density in a large region, shown in
Figure 5.9 c). The charge on the source side lowers the plateau of the hat-shaped
dependence which translates to change of the ID-VG slope in the linear region. The charge
on the drain side shifts ID saturation to lower values of VG. Charge on both sides combines
these effects.
For higher surface charge densities in a smaller region, the effects become more complex,
as shown in Figure 5.10 c). To the aforementioned effects of the charge one must add the
following. On the source side, it is VT shift to higher values of VG and reduced rate of
decline of gm in the region before saturation, hence shifting the saturation to higher values
of VG. On the drain side, it is blurring the transition from linear region to saturation, i.e. gm
starts to fall at lower values of VG, but the decline continues to higher values of VG than
without the trapped charge.
a) b)
Figure 5.10: Asymmetric charge distribution around
the gate. The impact of the trapped charge at the
source and drain sides of the gate is investigated
separately, for high charge density in a small region.
(a) ID-VG and (b) ID-VD characteristics and (c)
transconductance are compared.
c)
The best achieved fit of the pulsed I-V measurement, using uniform slabs of charge, is
shown in Figure 5.11. The charge density used in this simulation was asymmetrical, with
3×1012 cm-2 on the drain side and 2.5×1012 cm-2 on the source side of the gate, with
5.3 Investigation of the Impact of the Surface Electron Distribution 99
additional charge of 1×1012 cm-2 under the gate that provided for a threshold voltage shift.
The charge on the surface stretched 1m away from the gate. The blue ellipse in the figure
shows the region of I-V characteristics, where, for various examined combinations of
charge blocks on both sides of the gate and underneath it, the simulation results were
invariably divorced from the experimental results.
Examining the shape of the measured ID-VG characteristic suggested that there was trapped
charge on both the source and drain sides of the gate, with most of the charge on the drain
side. However, using uniform slabs of charge, a good fit to the experimental data was not
possible. These simulations indicated that to capture the current collapse behaviour, high
charge density close to the gate has to be present, whilst a low charge extended to
considerable distance away from the gate. This observation is consistent with the charge
distribution implied by the Poole-Frenkel transport mechanism observed in [97,127] and
modelled in [128].
Figure 5.11: The best fit of pulsed I-V characteristics using uniform blocks of trapped charge placed
asymmetrically around the gate. While achieving reasonable fit of I D-VD, as well as VT, and ON-current, the
transition between the linear regime and the ON-current in ID-VG (blue ellipse) remained problematic. This
appeared for all simulated values of VD, from 1V to 10V (not reported here).
5.4 Current Collapse Calibration: Exponential Charge Distribution 100
a) b)
Figure 5.12: a) Schematic illustration of the exponential charge distribution model, which, via the parameters
A, B and , allows for independent control of the sheet charge density at the gate edge QS(0), on the front of
the distribution QS(d) and the total charge QL trapped at the surface on either side of the gate. The method of
calculation of the parameters A, B and is described in the next subsection, 5.4.2, and given by equations
(5.3) and (5.4). b) Schematic illustration of the arrangement of the charge distribution in the device.
Exponential distribution on the source and drain sides of the gate and a constant sheet charge density under
the gate.
Results reported in this section are reported in [129]. Previous modelling efforts [128]
assumed a triangular charge distribution at the gate edge. For a more accurate modelling,
we have assumed here an empirical exponential function
(5.1) ( ) ( ) ,
for the surface charge distribution on both sides of the gate, where A, B and λ are fitting
parameters, which allowed to set the charge close to the gate, QS(0), very far from the gate,
QS(d), and the total charge on each side independently, QL, as shown in Figure 5.12 a).
Additionally, a constant charge density under the gate on the AlGaN/GaN interface was
assumed, as indicated in Figure 5.3.
In order to find the trapped charge distribution that would yield the best fit to the
experimental data using the above described model, we need to find the parameters A, B
5.4 Current Collapse Calibration: Exponential Charge Distribution 101
and λ, and the distance d to which the charge extended, for both sides of the gate, and the
value of the charge density under the gate QS,G, which, in total, results in nine parameters
to fit. To reduce the complexity of the problem, d was limited to dD = 1μm on the drain
side and dS = 0.5μm on the source side, since previous simulations have suggested that
there is larger charge on the drain side. These values were consistent with [71], where it
has been shown, that the electrons can migrate up to 1μm away from the gate. This reduces
the number of fitting parameters to seven. This arrangement is schematically illustrated in
Figure 5.12 b). The equation (5.1) allows us to calculate QS in a specific distance x from
the gate but, we need to control the variables Q(0), Q(d) and QL and calculate the fitting
parameters from these variables. First, from (5.1), we express the variables as
(5.2) a) ( )
b) ( ) ( )
c) ∫ ( ) ( ( ))
Now we need to express the parameters. First, we substitute (5.2) b) and then a) into c) and
rearrange the formula, to get λ, in equation (5.3) a). Then we substitute (5.2) a) into b) and
get the parameter A, in (5.3) b). Finally, rearranging (5.2) a) we get the parameter B, in
(5.3) c).
( ) ( )
(5.3) a)
( ) ( )
b) ( ( ) )
c) ( )
The obvious problem with this set of equations is that, to get A and B, we need to know λ
and, to calculate λ, we need to know B (or A). To solve this problem, we make a
preliminary assumption that, because the charge density drops down quickly as we move
away from the gate, ( ). We use this relationship to estimate the initial value of the
parameter λ and label it λ0,
5.4 Current Collapse Calibration: Exponential Charge Distribution 102
( ) ( )
(5.4) ( )
From this value we calculate A and then B and again λ. Then we update the value of λ until
the initial and final value is, within predetermined error, the same, i.e., | | .
1. Based on the insights gained from performing the simulations with uniform slabs of
charge in section 5.3.1, create a list of values pi,j for all seven fitting parameters,
where pi is a parameter, i = 1..7 and j = 1..Ni, Ni is the number of values of the
parameter pi in the list. The number of combinations is ∏ .
2. Read the grid file (produced by SSE (3.1.2) and read by SD (0)) to find out the
sequence of mesh points in the file, which is the same sequence as the one used in
the file specifying the doping (the doping file) in the device.
5. Using equation (5.1), calculate QS (and convert to charge concentration) for the
appropriate mesh points, defined as
6. Rewrite the appropriate mesh points in the doping file with the calculated values.
8. Have all combinations of the values of the parameters been selected? If yes,
continue. If not, go back to step 3.
9. List through all N results (ID-VG and ID-VD characteristics) and for each do the
following steps.
10. Linearly interpolate the I-V simulations to the values of V at which the current was
measured. E.g., we may have the value of Im at Vm from the measurement, yet Is1 at
Vs1 < Vm and Is2 at Vs2 > Vm from the simulation. In that case, calculate Is at Vm as
( ) . Do this for values of I at all measured values of V.
12. Choose the charge distribution with the lowest deviation as the one that fits the
experimental data best.
Table 5.1: List of values of the parameters used in the calibration of the exponential charge model. The top
table represents the first set of values, combinations of which were simulated. Based on the results from these
simulations, some values were swapped for new ones, shown in the bottom table. The new values are printed
in bold. QS (cm-2) is sheet charge and QL (cm-1) is total charge of the distribution.
The values used for the parameters mentioned in step 1 are listed in Table 5.1. Two sets of
simulations were performed. After the first run, the values that yielded the worst results
were dropped and substituted with new ones, printed in bold in the table. The values that
gave the best results from the second run of the simulations, which were better than the
results from the first run, are reported in the next subsection, 5.4.3.
a) b)
Figure 5.13: a) ID-VG and b) ID-VD calibration of
current collapse measurement, from class AB point,
at VD = 3V and at VG = 0V, respectively. Subfigure
c) shows the impact of the surface trapped charge on
transconductance, where the simulations without and
with the current collapse are represented by thin and
thick lines, respectively. The colours are consistent
with a). Figure 5.14 shows the corresponding surface
Figure 5.14: The surface charge distribution obtained by calibrating the pulsed I-V characteristics, shown in
Figure 5.13. The parameters of these distributions are summarized in Table 5.2.
The best fit to the ID-VD and ID-VG measurement achieved at the class AB operating point
is shown in Figure 5.13 left and right, respectively. The exponential distribution on both
sides of the gate for the best fit is shown in Figure 5.14 with the parameters given in
Table 5.2. Total charge on the drain side is more than five times higher than that on the
source side. The obtained distribution is consistent with the understanding that the current
collapse is due to charge injection from the gate corners and migration of the injected
charge away from the gate due to Poole-Frenkel emission mechanism. Space charge effects
in interplay with the trap activation energy and the corresponding „hopping mobility‟
determine the lateral distribution of the injected charge.
Table 5.2: Parameters of the exponential distribution that yielded the best agreement with experimental data,
as shown in Figure 5.13. QS (cm-2) is sheet charge and QL (cm-1) is total charge of the distribution at the
specified side of the gate. The surface electron distributions are visualized in Figure 5.14.
Position: Drain Side (dD = 1μm) Source Side (dS = 0.5μm) Gate
5.4.4 Accuracy
The accuracy of the calibration of the device I-V characteristics in the absence of current
collapse is discussed in section 3.4.3.3. Here we discuss the accuracy in the presence of
current collapse. In the first case, the calibration parameters are physical parameters with
known impact on the characteristics. Moreover, the range of possible values of these
parameters is known from literature. In the latter case, on the other hand, the calibration
parameter is the surface charge distribution, which, in principle, can be any function. We
have simplified the scale of this problem by making some assumptions about the
distribution, discussed in section 5.4.1, which made the calibration possible but in that we
have potentially sacrificed high accuracy of the simulated I-V characteristics. Therefore, in
the presence of the current collapse, one cannot expect to achieve the same accuracy as in
its absence.
a) b)
c) d)
Figure 5.15: a) The simulated ID-V characteristics during the current collapse (Figure 5.13 a) and b)
combined). The error of the calibration with respect to the measured characteristics (squares in Figure 5.13)
5.5 Device Degradation 107
(b) as a function of both VD and VG and separately, as a function of (c) VG and (d) VD.
Figure 5.15 shows the error of the collapsed drain current calibration. For the reported
values, the error is contained within 15%, with the majority of points being within 5%.
Subfigure c) reveals that the calibration is best for VG of -2 and -1V and subfigure d)
reveals systematic error that at lower drain voltage (VD < 4V) the drain current tends to be
overestimated while at higher drain voltage (VD > 6V) tends to be underestimated.
Performing further simulations using an optimisation procedure would most probably yield
more accurate calibration, but these results are sufficient to show that using asymmetrical
exponential surface charge distribution can lead to accurate description of the current
collapse phenomenon.
Figure 5.16: The lateral stress σx in the device at VG = -4V and VD = 25V as calculated from (4.4) within the
clamped model. The bottom subfigure offers a few cross-sections of the parameter. The value y in the top
subfigure indicates the distance from the interface while in the bottom subfigure it indicates the distance from
the surface. The gate is indicated as a pink line in the top subfigure and by the vertical lines in the bottom
subfigure.
The index i stands for an arbitrary mesh point of the device in the simulator. In both
models, defines the threshold stress, under which no charge is introduced at that mesh
point. In the constant charge density model, the parameter defines the charge density in
all mesh points, where the stress exceeds the threshold value. In the linear dependence
model, the parameter determines the increase rate of the trapped charge density as a
linear function of the lateral stress.
Figure 5.17: I-V simulations using the constant charge density model. The impact of the varying charge
density ρ0 was investigated, while the threshold stress σ0 was kept constant.
Figure 5.18: I-V characteristics for various values of the threshold stress and trapped charge density. The
third value in the legend is the total charge introduced in the device in each particular simulation.
From Figure 5.1 it is obvious that there was no VT shift after the device was degraded by
the stress test. From the simulations performed in the section 5.3 we know that VT shift is
caused by the charge trapped under the gate. When we combine these two pieces of
information, we can conclude that in calibrating the device degradation measurements we
need to avoid putting too much charge under the gate. Taking into account the 2D graph of
the lateral stress in the device in Figure 5.16, this gives us a lower estimate of the threshold
stress σ0 to be used in the calibration as 3.26GPa, preferably higher. We have to note here
that the maximum lateral stress in the device under the simulated bias conditions, which is
5.5 Device Degradation 110
right under the drain edge of the gate, 3.51GPa, is only 11% higher than the lateral stress
coming from the lattice mismatch between AlxGa1-xN and GaN for x = 0.28, 3.15GPa. The
highest stress here corresponds to the in-grown stress in the AlGaN layer for x = 0.31. If it
was the simple lateral stress that caused the defect formation and the following device
degradation, then a HEMT employing the Al0.31Ga0.69N/GaN heterostructure would be
degraded even without any stress test. This clearly is not the case. It is more probable that
what causes the defects are shear stresses and strains produced by the strong gradient of the
stress field. Therefore, the normal stress that is calculated by (4.4) and reported in
Figure 5.16 is taken only as a proxy to determine the region where the most defects will be
formed.
Figure 5.19: Simulated I-V characteristics using the linear dependence model for various values of the
parameters. Note that the total charge introduced in the device in the blue simulation is lower than in the
green simulation, yet the impact on the reduction of the drain current is stronger. Due to higher threshold
value, the region where the charge is introduced is smaller.
The measured I-V characteristics in this section, both before and after degradation, were
pulsed from an open-channel. The data labelled as fresh are the same as those labelled
open-channel in sections 5.3 and 5.4. Simulation results using the constant charge density
model are presented in Figure 5.17 and Figure 5.18. Figure 5.19 shows simulation results
using the linear dependence model. By using the lateral stress as a proxy to identify the
region where new traps are formed, it was unavoidable to introduce large density of charge
in the vicinity of the gate. Consistently with the investigation in section 5.3, such a charge
reduces the transconductance for VG just above VT. Since this is not what we see in the
experimental measurements of ID-VG characteristics of a degraded device, this effect
prevented us to find such parameters as to calibrate the measurements successfully.
Therefore we conclude that to reproduce the I-V characteristics of a degraded device it is
5.6 Summary 111
not sufficient to identify a region with high stress using the clamped model and to
introduce trapped charge in that region. The newly formed traps may form a gateway for
the electrons, an easier path to tunnel form the gate, to other, pre-existing traps.
5.6 Summary
Using the insights gained in this investigation, we have calibrated the measured I-V
characteristics with a manifested current collapse phenomenon (Figure 5.13). The trapped
charge distribution was found to be asymmetrical, with the majority of the charge residing
on the drain side, and non-uniform, with highest surface charge concentration close to the
gate while a comparably lower charge density extended out several gate lengths away from
the gate (Figure 5.14) in agreement with [128].
The estimated exponential shape of the surface charge proved to be sufficiently general to
reproduce the experimental current collapse measurements. Good agreement between our
simulation results and the experimental data was achieved, for a large range of simulated
voltages. This good agreement lends strong support to the virtual gate model and suggests
that non-linear transport of carriers injected from the gate via Poole-Frenkel emission [35]
is likely to be the explanation for the current collapse. Using this distribution as an ideal, in
Chapter 6 we will report on simulations of the electron leakage from the gate, by means of
Poole-Frenkel emission, to find the parameters that would reproduce the mentioned
distribution.
5.6 Summary 112
Using the clamped model and accounting for the converse piezoelectric effect, we have
determined a region in the device that undergoes high lateral stress during high power tests
at high voltages. We have used this information and introduced trapped charge into the
region to capture I-V characteristics of a degraded device. Within the selected model, it
proved impossible to fit the measurements.
6 Poole-Frenkel Electron Leakage
Mechanism
6.1 Introduction
We model the actual process of the electrons emitting from the gate, described in section
6.2, and “hopping” via surface trap states, described in section 6.3. In section 6.4 we list
the steps of the simulation. In section 6.5 we aim to calibrate the electron emission and
transport process that would lead to that distribution. In subsection 6.5.2.2 we compare the
distributions obtained by simulating the electron transport with the exponential distribution
found to reproduce the experimental I-V characteristics.
6.2 Electron Emission from the Gate 114
Figure 6.1: At high electric field E at the edge of the gate, electrons leak to the surface (red line). Then, due
to the strong E, they are transported away from the gate (blue arrows).
For temperatures above 250K, the electron leakage from the gate, which is a Schottky
contact, is described by Frenkel-Poole emission model [132]. In this model, the current
density is given as
( √ )
(6.1) [ ],
(6.2) ( ⁄ ) √ ,
where
√ , and .
The value of the parameter C isn‟t reported in the paper and cannot be recovered from the
graphs. In the above equation, the current density has dimensions A.m-2 (electrons that
6.2 Electron Emission from the Gate 115
cross a unit surface per unit time). However, we are interested in electron transport at the
surface and hence in the surface current density, j, the dimension of which is A.m-1
(electrons that cross a line of a unit length (the red line in Figure 6.1) per unit time).
Therefore, we will use the equation (6.2) only as a guideline and consider m and b to be
fitting parameters.
Since we do not know what the proper values of m and b for surface leakage current are,
we can tackle the problem from the other end. If we know, what surface current we expect
at two specific values of the electric field, we can calculate m and b. From equation (6.2)
we get
(6.3) ( )⁄ ( √ √ ), and
(6.4) √ ( ⁄ ),
where ji is the surface current density at the electric field Ei. Figure 6.2 shows the electric
field distribution at the surface of the device at a quiescent bias of VD,q = 25V and VG,q = -
4V. The electric field at the source and the drain edge of the gate is ES = 0.84MV/cm, and
ED = 3.51MV/cm, respectively. To estimate the surface current density at both edges of the
gate, jS and jD, we will use our exponential surface charge distribution calibration
(Figure 5.14) of the pulsed I-V characteristics. As reported in Table 5.2, the total amount
of the surface charge at the source and the drain side of the gate is QL,S = 6×107cm-1, and
QL,D = 3.5×108cm-1, respectively. The time between the pulses in this measurement was
tq = 1ms. This is the upper estimate of the leakage time, tleak, it takes for the electrons to
assume the final distribution. Thus, the average surface current densities throughout the
time between measurements tq, at the source and the drain edges, are ̅
, and ̅ , respectively. As the electrons leak to the surface, the
electric field at the gate edges is reduced and therefore the surface current densities are
reduced as well. To achieve an average ̅ , the electrons must start to leak with a higher
initial surface current density j,0, at the time of simulation t = 0. So, the reported estimates
of the average j are lower limits for the values used in the simulations. In principle, there
6.2 Electron Emission from the Gate 116
are two possible scenarios to be considered. One is that j,0 is only slightly higher than j
and reducing slowly, which results in tleak being close to tq. The other is that ,
resulting in most of the electrons being leaked in a short time, , and j dropping
dramatically thereafter with the electrons transporting and rearranging in the remaining
time, until the measurement. Obviously, there is no clear distinction between these two
scenarios and one naturally morphs to another. We will refer to these scenarios as low- and
high- density surface leakage current.
Figure 6.2: The magnitude of the x-component of the electric field at the surface of a HEMT device with
respect to the distance from the gate edge.
Figure 6.3: The dependence of the surface current density, j, on the x-component of the electric field for
various values of parameters m and b, calculated using equations (6.3) and (6.4). The values of E at the drain
(ED = 3.51 MV/cm) and source edge (ES = 0.84 MV/cm) of the gate, at VG = -4V and VD = 25V in the
simulated device, were taken for E1 and E2. The value of jD was set to 45 (solid lines) and
30 (dashed lines), the ratio jD/jS was set to 5 (blue), 10 (green), and 15 (red).
6.3 Electron Transport 117
Finally, Figure 6.3 demonstrates the dependence of the surface current density j on the
electric field for various selected values of jD and jD/jS. The graphs for higher jD look the
same, only with higher values on the y-axis.
Electron transport along the surface by the means of thermionic emission frequency can be
described by Eyring’s reaction rate model [133]
(6.5) ( ) ( ),
where is the electron emission frequency, h is the Planck constant, G is the activation
energy for the surface electron-hopping process and s is the spacing between the surface
trap locations. The velocity v of an electron is then defined as
(6.6)
[128] estimate that G = 0.25 0.3eV and s = 0.1 0.3nm. Figure 6.4 and Figure 6.5 show
the electron emission frequency and electron velocity, given by equations (6.5) and (6.6),
for this range of values, respectively.
The above mentioned range of values of G and s translates to a wide range of electron
velocities. In the linear region, the variation between the highest (G = 0.25eV, s = 0.3nm)
and lowest (G = 0.3eV, s = 0.1nm) velocities is more than sixty-fold. This results in a
great uncertainty in the simulations. Table 6.1 shows the combinations of values of these
two parameters used in the simulations to represent the wide range of electron velocities in
this model. We also report on the time the electric field was frozen for, tf. The higher was
the velocity, the shorter was tf.
6.3 Electron Transport 118
Figure 6.4: The dependence of the emission frequency, with which an electron tunnels from a surface trap to
the next trap, on the electric field, according to the Poole-Frenkel transport model, expressed by equation
(6.5).
Figure 6.5: The velocity of an electron travelling across a HEMT surface travelling from trap to trap, as given
by equation (6.6).
Table 6.1: The values of physical parameters, that enter the Poole-Frenkel transport model, used in the
simulations.
We have performed self-consistent simulations taking into account the impact of the
leaking electron distribution on the electric field at the gate edge and device surface. The
steps in the simulation procedure are as follows:
2. Extract the electric field at the surface from the simulation. The electric field is
frozen for time tf.
3. Inject jS × tstep electrons at both edges of the gate, according to the equation (6.2)
and the local electric field, where tstep << tf is a time step, and tf = n × tstep, where n
is an integer.
4. Move the surface electrons by v × tstep, where v is the electron velocity given by the
equations (6.5) and (6.6), shown in Figure 6.5.
6. According to their positions, assign the electrons the appropriate mesh points of the
device in the simulator.
a) b)
c) d)
Figure 6.6: Profiles of the electron density on the device surface at the drain side of the gate. The gate edge is
positioned at x = 1.25m. All four graphs show four different values of initial surface currents on the drain
edge of the gate, jD,0, namely 50, 45, 40 and 35 e-m-1s-1, represented by thick solid, dotted, thin solid and
dashed lines, respectively. The ratio between initial surface currents at the drain and source sides is in all
cases jS,0/jD,0 = 5. The top row shows the distribution at two specific moments, using (a) slow and (b)
moderate velocity profiles. The bottom row compares the distributions that result from different velocity
profiles, at a specific time of the simulations. c) Slow and moderate at 400s, and (d) slow, moderate and fast
at 160s.
Figure 6.6 shows the electron distributions at different moments, resulting from different
initial surface electron density and different electron velocity profiles, defined in Table 6.1.
The unsurprising result of comparing the velocity profiles is that, the higher the velocity,
the further the electrons travel away from the gate. However, it is interesting to note that
with higher velocities, the electron distribution is not only stretched, but skewed as well.
Increasing the velocity near the gate edge lowers the electron density there
disproportionately. This phenomenon can be explained as follows. First, for lower velocity,
at a specific time, the electrons are closer to the gate. This leads to reduction (in absolute
6.5 Simulation Results 121
terms) of the electric field closer to the gate, and hence the reduction of velocity and
further accumulation of electrons there, in a positive feedback. Lower velocity profile will
lead to higher electron concentration closer to the gate.
a) b)
Figure 6.7: a) The electric field distribution using slow and moderate velocity profiles after 160s and 80s,
respectively. The black line is the electric field before the electron leakage. Higher leakage current and hence
higher electron density (solid lines) leads to higher electric field and therefore higher velocity at the front of
the electron distribution. This results in the electrons getting farther from the gate. The colouring corresponds
to the Figure 6.6 c). b) The electric field (arrows) before (top) the leakage and the electric field (blue and
green) associated with the leaked electrons (bumps) and its effect on the total electric field (red arrows with
blue and green outlining). Higher electron concentration increases the electric field and velocity at the front
of the stream of electrons and reduces it at the end close to the gate. The vertical black line represents the
position of the gate edge. Note: The bump on the graph of Ex in the figure on the left is due to an abrupt
change in mesh spacing in the simulator.
Another observation is the fact that increasing the electron surface leakage current not only
provides more electrons to the surface, but also stretches the electron distribution in both
directions, towards and away from the gate. To understand why, we need to turn our
attention to the scheme in Figure 6.7 b). The thick black line represents the gate edge, and
the arrows represent the strength and direction of the electric field. The top row shows the
situation before any electrons leak to the surface. There is a very high electric field at the
gate edge (red arrows), which decreases rapidly with distance (as shown in Figure 6.2).
Then the electrons start to leak to the surface. We consider higher and lower jD,0, which
results to different amount of electrons, represented by blue and green bumps, respectively.
6.5 Simulation Results 122
The electric field associated with the electrons (blue and green arrows) is proportional to
their concentration. In the case with more electrons, the resulting electric field (red arrows
with blue outlining) is lower closer to the gate, and higher at the other end. From this
follows that, the electrons closer to the gate will move slower than in the case with fewer
electrons (red arrows with green outlining). This is illustrated in Figure 6.7 a). Higher jD,0
results in higher electron density and higher negative electric field at the front of the
electron distribution, which in turn means higher velocity.
a) b)
Figure 6.8: The electron distribution on (a) the source and (b) drain sides of the gate after 160s (solid) and
400s (dashed). The simulations were done for all three above mentioned velocity profiles, slow (blue),
moderate (red) and fast (green). b) (drain side) also shows the impact of changing the jD,0/jS,0 ratio (light
coloured lines). The effect is stronger for slower velocities. Using the fast velocity profile, the electrons on
the drain side crossed almost 1m in just 160s. The gate is positioned between x=1m and x=1.25m.
The electron sheet density on both sides of the gate, for various velocity profiles, is shown
in Figure 6.8. In the subfigure b), showing the drain side of the gate, two ratios of initial
current on the drain and source sides are compared, jD,0/jS,0 = 5, 10. This has impact on the
dependence of the surface current leakage on the electric field, as shown in Figure 6.3.
When the electrons are closer to the gate, due to the lower velocity, they reduce the electric
field at the gate edge more. Therefore, the effect is more pronounced for lower electron
velocity. On the source side, the effect is negligible for the two simulated ratios.
6.5 Simulation Results 123
a) b)
Figure 6.9: The electric field at the drain (a) and source (b) edges of the gate. Red lines show simulations
with same electron leakage parameters (jD,0, jD,0/jS,0) and different velocity profiles, given by different
combinations of ΔG and s. Solid lines show simulations with slow electron velocity, the same jD,0/jS,0 ratio,
but varying initial surface current density jD. On the drain side, pink and orange dashed lines represent
simulations with the same jD as the red line, but varying jD,0/jS,0 ratio.
a) b)
Figure 6.10: Due to the electric field at the edge of the gate, the electrons tunnel from the gate to the device
surface. a) Electron surface current density on the drain edge jD given by the Frenkel-Poole emission model.
b) The total charge leaked to the surface. The colour coding is the same as in Figure 6.9, except for the fast
velocity profile in b), where the dotted red line is replaced with solid black, to make the graph more readable.
The temporal evolution of the x-component of the electric field, Ex (at the drain and source
edges of the gate), the surface leakage current density jD and the total leaked charge at the
surface (on the drain side of the gate), which is an integral of jD, ( ) ∫ ( ) , are
shown in Figure 6.9 and Figure 6.10, respectively. The red lines show results from
simulations with the same leakage parameters, jD,0 = 50 e-m-1s-1 and jD,0/jS,0 = 5, and
different velocity profiles (Table 6.1). The higher are the velocities, the further away the
electrons move from the gate and the less the electric field at the gate edge is decreased.
6.5 Simulation Results 124
From this follows that, lower velocities reduce the leakage current density faster and, as a
result, fewer electrons leak to the surface. Solid lines show the impact of varying jD,0,
keeping the velocity profile (slow), ΔG = 0.3eV and s = 0.1nm, and the ratio of the initial
surface current densities on the drain and source side, jD,0/jS,0 = 5, constant. Unsurprisingly,
the higher is the jD,0, the stronger is the effect on the electric field and all related quantities.
Let us note that, even the relative effect on jD is stronger. After 800s of simulation,
jD = 0.65 (0.74) jD,0, for jD,0 = 50 (35) e-m-1s-1. Finally, the impact of changing the
electron leakage dependence (Figure 6.3) on the electric field on Ex, jD and QL at the drain
edge of the gate, by setting jD,0/jS,0 to 7 and 10, is represented by the pink and orange
dashed lines, respectively. Since this parameter affects the current density jD at lower
electric fields, increasing the jD,0/jS,0 ratio reduces the total leaked charge QL and the
reduction of the electric field at the gate edge is less pronounced.
To address these issues, simulations with high jD,0 (200-2000 e-m-1s-1) employing the
slow and moderate velocity profiles and very high jD,0 (5000 e-m-1s-1 and more)
employing only slow velocity profile, were performed in 6.5.2. Additionally, the impact of
the values used for the ratio jD,0/jS,0 (5, 7 and 10) on the electron distributions (Figure 6.8 b)
was negligible. Therefore, in the simulations that follow, we used the values 5 and 20 for
the jD,0/jS,0 ratios.
Table 6.2 lists all simulations that were performed in this section, along with the colour
codes used in the following figures. The colour coding is not kept when comparing
simulations with the same jD,0. The values in the table report on the time of the simulations
the electric field was kept constant (frozen). A simulation of the device, and hence the
recalculation of the electric field, takes approx. 30-40 minutes. So, choosing longer tf
allows us to simulate the electron leakage for a longer time of the process for a given
simulation time, or, simulation of a given real world time takes less simulation time. On
the other hand, with the electrons changing their positions, the electric field distribution
changes as well and therefore the electron leakage and transport are also changed. Not
updating the electric field often enough leads to huge errors. Obviously, there is a balance
to be struck. This issue is addressed in section 0, where the simulations with the same
parameters, but different tf are compared, as noted in Table 6.2.
Table 6.2: List of simulations analysed in this section. tf is the time step, for which the electric field was
frozen in each simulation. Where there are two numbers, two simulations were performed, to analyse the
impact of the time step on the evolution of the electric field and the resulting electron distribution. The
colours of the lines represent the colours used in the following figures for the respective simulations.
jD,0 ( )
tf(rozen) (s)
200 500 1,000 2,000 5,000 10,000 20,000 50,000 75,000
jD,0/jS,0 ΔG (eV)
5 2 0.8 0.4 0.4 0.4 0.2 0.1 0.2/0.1 -
0.3
20 2 0.8 0.4 0.4 0.4 0.4 0.1 0.2/0.04 0.4/0.02
5 2 0.8 0.4 0.4 - - - - -
0.25
20 2 0.8 0.4 0.4 - - - - -
6.5 Simulation Results 126
In reporting the simulation results of quantities such as the x-component of the electric
field Ex at the source/drain edge of the gate, the surface electron leakage current to the
source/drain side jS/jD, total amount of electrons leaked to the surface QL of the
source/drain side of the gate, and finally the electron distributions themselves, we
concentrate on the drain side, since the leakage is in principle the same. To capture the DC-
RF dispersion, we are mostly interested in the final distribution of the electrons in the
moment of the pulsed measurement, and whether such a distribution is actually assumed or
not. To achieve a final distribution, one of two scenarios must occur. Either, the leakage of
electrons and their velocity drop to very low levels so that the distribution is more or less
fixed for the time of the measurement and all subsequent measurements, or the leaked
electrons are compensated by the flow of electrons that are transported away by the
Frenkel-Poole mechanism. To put it in other words, either, the leakage and subsequent
flow of the electrons, due to low values of the electric field, stops or significantly slows
down, or, the leakage and transfer of electrons reach a steady state at which the overall
distribution does not change, while the surface electrons continue to move from trap to
trap. If neither of those can be achieved, the whole concept of using a single surface
electron distribution to calibrate the pulsed I-V characteristics is questionable.
6.5.2.1 The electric field at the Gate Edges and Emission of the Electrons
from the Gate to the Surface of the Device
An essential part, in search for the answer to the question of the stability of the surface
electron distribution, is obtaining Ex at both of the gate edges, which, via the Poole-Frenkel
emission model, determines jD,0 and jS,0. Ex is reported in Figure 6.11, Figure 6.13 and
Figure 6.15 for lower and higher jD,0 on the drain edge and all values of jS,0 on the source
edge of the gate, respectively. The corresponding jD and jS are reported in Figure 6.12,
Figure 6.14 and Figure 6.16. We can make some observations based on this set of figures.
The impact of the ratio jD,0/jS,0 (5 vs. 20) and of the velocity profile, represented by the
parameter ΔG (0.25eV vs. 0.3eV), as well as the jD,0 parameter, are compared. Firstly,
higher jD,0 means more electrons burst to the surface initially, which leads to faster
reduction of the electric field. This in turn reduces jD and jS faster. Not only relatively, as a
portion of jD,0 and jS,0, but, as may be seen on the right of the figures showing jD and jS, also
in absolute terms. This effect is strongly amplified with the increase of jD,0 and jS,0. For the
6.5 Simulation Results 127
Figure 6.11: Ex at the drain edge of the gate simulated for various values of jD,0, using two different velocity
profiles and two different electron leakage profiles, with one combination missing, for the sake of clarity of
the figure. This is a follow-up to Figure 6.9 left, for simulations with higher jD,0, with the highest jD,0 from
that figure reprinted in this one with grey lines. Follow-up to this figure with even higher jD,0 is Figure 6.13.
Increasing jD,0 further reduces the electric field even faster. For jD,0 = 2,000 e-m-1s-1, the electric field at
the gate edge drops to half its initial value in less than 6 s. This has a huge impact on the development of the
electron distribution. The corresponding electron distributions are presented in Figure 6.17: solid lines
(ΔG = 0.3eV, jD,0/jS,0 = 5) – top left; dashed lines (ΔG = 0.3eV, jD,0/jS,0 = 20) – top right; dotted lines
(ΔG = 0.25eV, jD,0/jS,0 = 5) – bottom left; unreported here (ΔG = 0.25eV, jD,0/jS,0 = 20) – bottom right.
Figure 6.20 right compares the distributions (top) and jD (bottom) simulated with the same value of
jD,0 = 2,000 e-m-1s-1 (red lines here). The corresponding jD is shown in Figure 6.12.
a) b)
Figure 6.12: The dependence of the temporal evolution of jD on jD,0, on the emission model parameters and
on the transport velocity of the electrons. Higher jD,0 results in faster reduction of jD both relative to jD,0 (left)
and in absolute values (right). QL, which is the integral of jD, for ΔG = 0.3eV is reported in Figure 6.22 and
Figure 6.23 for jD,0/jS,0 = 5 (solid lines) and jD,0/jS,0 = 20 (dashed lines), respectively. The corresponding Ex is
shown in Figure 6.11, all other parameters as described therein.
6.5 Simulation Results 128
Figure 6.13: Ex at the drain edge of the gate, for jD,0 = 5,000 e-m-1s-1 and more, which is a follow-up
figure to Figure 6.11 (with one of the simulations shown in that figure reprinted here). As expected, further
increase in jD,0 speeds up the reduction of Ex at the gate edge and hence the electron tunnelling to the device
surface. For the parameters jD,0 = 20,000 e-m-1s-1 and jD,0/jS,0 = 5, Ex drops to half its initial value in
0.47s, and for jD,0 = 50,000 e-m-1s-1 and jD,0/jS,0 = 20 in 0.25s. Some of the corresponding electron
distributions are reported in Figure 6.18 (jD,0/jS,0 = 5), Figure 6.19 (jD,0/jS,0 = 20) and Figure 6.21
(jD,0 = 20,000 e-m-1s-1). The corresponding jD is shown in Figure 6.14.
a) b)
Figure 6.14: Further increase of jD,0 causes further acceleration of the reduction of jD (left) which, for high jD,0
quickly drops below the jD of the lower jD,0 (right). The higher is difference between jD,0 of two simulations,
the faster this happens (right). QL is shown in Figure 6.22 and Figure 6.23 for jD,0/jS,0 = 5 (solid lines) and
jD,0/jS,0 = 20 (dashed lines), respectively. The corresponding Ex is shown in Figure 6.13, all other parameters
as described therein. Note: the slight oscillation in the simulation of jD for jD,0 = 20,000 e-m-1s-1 and
jD,0/jS,0 = 5 (solid black line) is caused by the fact that the time step was insufficiently short. This will be
expanded on in the next section (0).
6.5 Simulation Results 129
parameters jD,0 = 50 e-m-1s-1 (jD,0/jS,0 = 5) and lower, Ex at the gate edge did not even
approach half its initial value in the simulated time of 800s, as reported in the previous
section in Figure 6.9, and included for comparison in Figure 6.11. In contrast, for
jD,0 = 2,000 e-m-1s-1 (jD,0/jS,0 = 5) Ex reached half its initial value in just under 6s, for
jD,0 = 20,000 e-m-1s-1 (jD,0/jS,0 = 5) in 0.47s and for jD,0 = 50,000 e-m-1s-1
(jD,0/jS,0 = 20) in 0.25s. Therefore, to shut off the leakage of electrons from the gate, the
value of jD,0 must be even higher than those used in the reported simulations.
Another observation from these figures is that higher-velocity profile impedes shutting-off
the electron emission from the gate. This is due to the fact that higher velocities move the
electrons faster away from the gate and hence they have lower impact on the Ex reduction.
Similarly, higher value of the jD,0/jS,0 ratio has a similar effect on Ex. To conclude, high jD,0,
low jD,0/jS,0 and low velocity profile are conductive in shutting off the electron leakage
from the gate.
It has to be noted that two of the reported simulation results here show clear signs of
insufficiently short tf with respect to the chosen parameters. It shows as oscillations in the
time dependence of Ex and/or j. These are the jD,0 = 20,000 e-m-1s-1, jD,0/jS,0 = 5 and
tf = 0.1s, reported as solid black lines in Figure 6.14 and the jS,0 = 1,000 e-m-1s-1,
jD,0/jS,0 = 5, ΔG = 0.3eV and tf = 0.4s, reported as solid light green line in Figure 6.15 and
Figure 6.16. However, since this behaviour follows the general dependence and trend set
by other results, we believe that apart from the oscillations, they approximate the correct
result reasonably well.
6.5 Simulation Results 130
Figure 6.15: Ex at the source edge of the gate. The corresponding distribution and jS for jS,0 = 100 e-m-1s-1
(orange lines), compared with jD,0/jS,0 = 20 (not shown in this figure) is shown in Figure 6.20, top and bottom
left, respectively. The jS corresponding to this figure are shown in Figure 6.16.
a) b)
Figure 6.16: jS for various jS,0 and various velocity profiles. QL for ΔG = 0.3eV is reported in Figure 6.24 left;
the corresponding Ex in Figure 6.15; and other parameters as described therein. The cause of the oscillations
for jS,0 = 1,000 e-m-1s-1 (solid light green line) is a too large time step.
After investigating the electron leakage to the surface of the device, let us turn our
attention to the actual electron distribution, and how it evolves with time, to see, whether
or not, the distribution comes close to the one obtained by calibrating the pulsed I-V
6.5 Simulation Results 131
characteristics, shown in Figure 5.14. Let us call it the ideal distribution to refer to later in
the text. Figure 6.17 shows the distributions on the drain side of the gate at two time points,
at 20s and 100s, for moderate initial surface electron leakage currents, jD,0 = 200 –
2,000 e-m-1s-1. The four subfigures compare the impact of the velocity profiles, defined
by the parameter ΔG, and impact of the emission model parameter jD,0/jS,0. Figure 6.18 and
Figure 6.19 show the distributions for high values of leakage current, jD,0 = 2,000 –
50,000 e-m-1s-1, each using 5 and 20 as values of the parameter jD,0/jS,0, respectively,
this time only for the slow velocity profile. Figure 6.20 compares the impact of the velocity
profile and the ratio jD,0/jS,0 on the distributions and jD on both sides of the gate in one
figure, keeping jD,0 constant, at 2,000 e-m-1s-1. Figure 6.21 shows the impact of jD,0/jS,0
on the distribution on the drain side, for the slow velocity profile and for jD,0 = 20,000 e-
m-1s-1.
Figure 6.17: The electron distributions at the drain side of the gate for four different values of jD,0 at two time
points, 20s (dotted lines) and 100s (solid lines). Figures shows simulation results for slow (top) and
moderate (bottom) velocity profiles, and two electron emission profiles, jD,0/jS,0 = 5 (left) and jD,0/jS,0 = 20
(right). The distributions from simulations for jD,0 = 2,000 e-m-1s-1 are compared in Figure 6.20, right. Ex
and jD for all simulations, except the one on bottom right, are reported in Figure 6.11 and Figure 6.12; QL for
top left (ΔG = 0.3eV, jD,0/jS,0 = 5) and top right (ΔG = 0.3eV, jD,0/jS,0 = 5) are shown in Figure 6.22 and
Figure 6.23, respectively.
6.5 Simulation Results 132
The overall trend, when increasing jD,0, is clear. For higher jD,0, the electrons tend to
accumulate close to the gate and also spread further away from the gate. However, the
latter effect diminishes with increasing the jD,0. Lower velocity and lower jD,0/jS,0 give
distributions closer to the ideal, which is clear especially in Figure 6.20. For the lower
velocity, the electrons reach only to 300m away from the gate in 100s, which is
reasonable, since tq = 1ms, and there is enough time to cover 1m, maybe more.
Figure 6.18: A follow-up to Figure 6.17 top left, the electron distributions for three different values of jD,0 at
three time points, 2s, 10s and 30s. In spite of grand differences in distributions shortly after the start of
the leakage, at 2s (dotted lines), due to unequal drop in Ex and hence jD, the distributions converge to
roughly the same “shape” later, at 30s (solid lines). One of two main differences among the distributions is
the maximal distance the electrons reached, which changes less and less with higher values of jD,0 (compare
with Figure 6.17). The other is that with higher jD,0, the electron density at the gate edge is higher (right). The
corresponding Ex and jD are reported as solid lines in Figure 6.13 and Figure 6.14, respectively; the
corresponding QL is shown in Figure 6.22.
6.5 Simulation Results 133
Figure 6.19: A follow-up to Figure 6.17 top right, the electron distributions for very high jD. As in true for
simulations in Figure 6.18 (jD,0/jS,0 = 20), higher jD means more electrons at the gate edge and the front of the
distribution is further away, although the latter is less significant for very high jD. Otherwise, the resulting
distributions are similar. The corresponding Ex and jD are reported as dashed lines in Figure 6.13 and
Figure 6.14, respectively; the corresponding QL is shown in Figure 6.23.
Figure 6.20: The electron distributions (top) and jS as a function of time (bottom) for the source (left) and
drain (right) sides of the gate. This figure compares simulations with the same jD,0 (jS,0), for two different
velocity profiles (given by the parameter ΔG) and two different electron emission parameters (jD,0/jS,0).
Slower velocity and lower jD,0/jS,0 both mean more electrons close to the gate and hence a distribution closer
to the “ideal”. On the drain side, the corresponding Ex and jD are reported as red lines in Figure 6.11,
Figure 6.12, respectively. Blue and green lines in this figure are results of simulations with the same
parameters for both the source and the drain sides. The distributions (top) are a higher jD follow-up to
Figure 6.8.
6.5 Simulation Results 134
Figure 6.21: While Figure 6.18 and Figure 6.19 compare the electron distributions for varying jD,0, while
keeping jD,0/jS,0 constant, this figure compares the impact of varying jD,0/jS,0 at jD,0 = 20,000 e-m-1s-1. It is a
higher jD follow-up to Figure 6.20 top right for ΔG = 0.3eV. As, keeping the previous results in mind, one
would expect, the difference between these simulations reduces with time and the main difference is the
electron density close to the gate. The corresponding Ex and jD are shown in Figure 6.13 and Figure 6.14,
respectively.
Figure 6.22: The total amount of the electrons leaked to the drain side of the gate for various jD,0 at
jD,0/jS,0 = 5, for slow velocity profile. The higher is the initial leakage, the more electrons leak to the surface.
However, since jD falls rapidly for simulations with high jD,0 (Figure 6.12 right and Figure 6.13 right), the
increase of QL , after the initial burst, reduces faster for higher jD,0. This is well demonstrated in the figure on
the top right. The corresponding electron distributions are reported in Figure 6.17 top right and Figure 6.18.
6.5 Simulation Results 135
Another important parameter of the simulations is the total amount of electrons leaked to
both of the sides of the surface, QL, since we have an estimate of how many electrons we
should expect, reported in Table 5.2 in section 5.4. Figure 6.22 and Figure 6.23 show QL
leaked to the drain side of the gate for jD,0/jS,0 = 5 and jD,0/jS,0 = 20, respectively.
Figure 6.24 shows the same on the source side. It is encouraging, that for the wide range of
values of jD,0, which covers more than two orders of magnitude, QL is in the vicinity of the
expected values, even though, for longer simulation times, QL exceeds the expected value.
For all simulated values of jD,0 and simulated times, it is true that higher jD,0 reach higher
QL at a specific time. However, as may be seen in Figure 6.22 top right, the higher the jD,0
is, after the initial burst, the slower the QL grows, so, in a finite time, higher jD,0 may result
in lower QL. To see that, we would need to run the simulations for longer and for even
higher jD,0.
Figure 6.23: QL for various jD,0 at jD,0/jS,0 = 20, for slow velocity profile. The corresponding electron
distributions are reported in Figure 6.17 top left and Figure 6.19.
Figure 6.24: The electrons leaked to the surface side of the gate for jD,0/jS,0 = 5 (left) and jD,0/jS,0 = 20 (right).
6.5 Simulation Results 136
We have investigated the electron emission from the gate and subsequent transport at the
surface for a wide range of values of the initial surface current density. Here we estimate
the plausibility of the used values. The highest value used in our simulations, as reported in
Table 6.2, was 75‟000 e-m-1s-1, which translates to 1.2×10-8Am-1. The electrons leak
to the surface at the quiescent bias VG = -4V, VD = 25V, at which the drain current in this
device was measured to be 180mA, which is 3.6×10-4Am-1, considering the 4×125m
width of the device. The charge in the channel is confined to a very narrow region next to
the interface. As an upper limit, we can take 1.5nm, at which the charge density falls
approximately tenfold, resulting in average current density j = 0.24Am-2. It is not clear
what is the thickness of the path through which the surface electrons propagate. As an
approximation, one could take the lattice constant c, which is the height of the wurtzite
crystal, equal to approximately 0.5nm, as reported in Table 2.1. Then the current density of
surface leakage will be j = 2.4×10-4Am-2. This is three orders of magnitude less than the
current density in the channel. This constitutes a reasonable limit to the value of the surface
current density.
6.5.3 The impact of the Time Step on the Surface Charge and
Leakage Current
Figure 6.25 shows jS (left) and jD (right) for simulations with very high jD,0. Different tf are
compared. tf is shortened two-fold, five-fold and twenty-fold, in the top, centre and bottom
subfigure. It shows that the higher is the ratio between the values of tf, the change and
precision enhancement is increased. Too large values of tf result in oscillations of j between
values that are alternately too low or too high. In the first case, the oscillations are reduced.
The second case is a good example how reducing tf can lead to increasing the precision of
the simulation, while in the third case, the larger tf is clearly inapplicable, but reducing the
tf twenty-fold prevents jD to swing to negative values.
6.5 Simulation Results 137
Figure 6.25: The impact of the time step on the simulations, here exemplified using the physical parameters jS
(left) and jD (right). The parameters used for these simulations are as follows. Top: jD,0 = 50,000 e-m-1s-1,
jD,0/jS,0 = 5; Centre: jD,0 = 50,000 e-m-1s-1, jD,0/jS,0 = 20; Bottom: jD,0 = 75,000 e-m-1s-1, jD,0/jS,0 = 20.
The difference in the time the electric field was kept constant (frozen) tf was also different in these
simulations. The decrease in tf was two-fold (top), five-fold (centre) and twenty-fold (bottom). The higher is
the difference in tf, the more the simulation is changed.
Fluctuations at the front of the surface charge distribution can be observed in the
simulation results, e.g. in the Figure 6.6, Figure 6.17 and Figure 6.20. This is not a real
effect, it is a residue of too large a time step for which the electric field is kept constant.
The reason behind this effect following. In the absence of trapped electrons at the surface,
the lateral electric field Ex is highest at the gate edge and monotonously declines towards
the other contacts. Yet, when the electrons leak to the surface, they modify Ex and, for
concentrations high enough, due to the repelling force of the electrons on other electrons,
6.6 Summary 138
Ex increases and then falls sharply. This is demonstrated in Figure 6.7 a) by the orange
lines, and explained in the caption more thoroughly. This peak and sharp decline in the
electric field translates to higher and lower velocities respectively. In a real world scenario,
the position of the peak moves as the electrons move away from the gate. Yet when the
electric field is held constant for a specific time step, some electrons leak behind the peak
and their velocity is first artificially increased at the peak position and later decreased.
This, in turn, leads to artificial decrease of the concentration at the peak position and
accumulation of electrons thereafter. The created fluctuation in the electron concentration
translates into another fluctuation of the electric field and in positive feedback more peaks
and valleys in the surface charge distribution are introduced. Reducing the time step
reduces this unwanted effect due to more frequent recalculation of and hence more realistic
electric field. However, this would result in an unwanted increase in simulation time.
Therefore, we sacrifice some precision in prospect of shorter simulation time.
6.6 Summary
We found that, for high values of the initial electron surface leakage jD,0 and low velocity
profile, the assumed electron distribution resembles the ideal the most, in that it follows the
pattern, a lot of electrons at the gate, less electrons further away, stretching for several gate
lengths. In simulations with high value of jD,0, the electron leakage drops fast to low levels
and hence has the potential to fix the amount of leaked electrons in some time. However,
for none of the simulated parameters, the velocity of the electrons seems to drop enough
for the electrons to effectively stop moving and even the promising distributions, i.e. those
with high electron density close to the gate, lose those electrons in further electron
transport away from the gate via traps. To determine whether it is possible to achieve a
quasi-final electron distribution, it would be beneficial to run the simulation for longer
“real” time, which would require longer simulation time, and with higher jD,0, which would
require shorter tf, also resulting in longer simulation time. In Figure 6.25, we have
demonstrated that, to simulate the electron leakage for jD,0 = 75,000 e-m-1s-1, one has to
consider tf no longer than 0.02s. To simulate at least one interval between the
measurements, tq = 1ms, we would need 50,000 cycles, each taking approximately half an
hour, resulting in enormous 25,000 hours, i.e., almost three years.
6.6 Summary 139
The charge close to the gate and the electric field come to equilibrium and the charge flow
(leakage current density) is stabilized and never falls to zero. Moreover, the field further
away from the gate is non-zero as well – that means that the electrons never cease to move
and hence it is probably not possible to arrive at a “final” distribution.
7 Conclusions
The aim of this thesis was to investigate some of the processes determining the degradation
and failure of AlGaN/GaN HEMTs, namely DC-RF dispersion, a subclass of current
collapse phenomena, and device degradation, using numerical simulations. The
commercial simulation platform Sentaurus from Synopsys and a set of scripts to
manipulate and automate the device simulations were utilised and developed to perform
this task. Both the current collapse and the device degradation are trap-related phenomena.
DC-RF dispersion is associated with charge leaked to and trapping mainly at the surface of
the device, in already existing traps, forming a „virtual gate‟ [116,34]. The device
degradation is a process that leads to creation of new defects and dislocations and hence
new traps in the device and has a similar, even though not identical, permanent effect on
the I-V characteristics as the current collapse. One of the recent hypotheses for the
mechanism responsible for the device degradation is defect generation due to electric field
induced stress [29] in the vicinity of gate edges, particularly on the drain side, with the
converse piezoelectric effect as the underlying physical phenomenon. Another
consequence of the converse piezoelectric effect is related to the additional strain in the
material that modifies the spatial distribution of the piezoelectric polarization, the bound
charge and the transistor characteristics. The current collapse is associated with mobile
charge leakage along the surface via Poole-Frenkel emission [36,35].
determine suitability of a material for fabrication of such devices are the saturation velocity
and the breakdown electric field, as given by the Johnson‟s figure of merit (JM) [3]. The
comparison of the values of JM for various materials makes GaN stand out as the material
of choice. The two classes of GaN devices with high application potential, the
optoelectronic devices [14,15,16] and the high-frequency transistors [13,2], were
introduced, later of which is the focus of this work.
Chapter 2 started by discussing the material parameters of III-Ns with focus on the
properties that make this class of materials distinct from the conventional III-Vs, such as
the wurtzite crystal structure with the consequence of spontaneous and piezoelectric
polarization. This results in a bound charge at a III-N heterostructure interface, which gives
rise to a large 2DEG density in the channel without the need for doping. Relation between
the electrical properties, i.e. the electric field and piezoelectric polarization, and
mechanical properties, i.e. the stress and strain, was discussed. The related direct and
converse piezoelectric effects were introduced with the implication of the electric field
induced strain and stress, utilized in Chapters 4 and 5 in the investigation of the gate
dependent bound charge and device degradation, respectively. Then, the principle of
operation of AlGaN/GaN HEMTs, the device, which is in the focus of this thesis, was
discussed and an overview of key challenges in current GaN technology was provided.
Voltage applied to the gate modifies the electric field at the edges of the gate and in the
channel under the gate significantly. This leads to strain, induced via the converse
piezoelectric effect, being dependent on the gate voltage. From this follows that the
piezoelectric polarization and hence the bound charge, induced in the bulk and modified at
the heterojunction interface, will vary with the applied voltage too. This phenomenon was
investigated in Chapter 4. First, the theory behind the converse piezoelectric effect was
elaborated. Because the computation of the stress and strain distribution from the electric
field in the device is a complex electro-mechanical problem that goes beyond the scope of
this thesis, a simplifying „clamped model‟ [121,30] was adopted, which decouples the x
and y components of the electric field from the mechanical properties of the crystal and
also ignores the impact of mechanical stresses of the surrounding material on the local
strain. In this chapter, we have, for the first time, evaluated the magnitude and the
importance of this effect. Even without the impact of the gate voltage, the coupling of the
electric field to the piezoelectric polarization decreased the difference in polarizations of
the AlGaN and GaN layers and hence decreased the polarization induced sheet charge that
lead to reduction of the drain curren, although this effect is moderate, leading to less than
1% current change at VG ≈ 0V. However, negatively increasing the gate voltage amplified
this effect and the drain current reduction becomes more pronounced, leading to the
threshold voltage shift. The current reduction was found to be virtually independent of the
drain voltage.
Since both the current collapse and device degradation are trap related phenomena,
Chapter 5 first investigated the impact of electrons, trapped at the surface of the device and
under its gate, on the I-V characteristics. In order to gain insight, uniform distributions of
charge on either side of the gate were used both with symmetric and asymmetric charge
distributions in respect to the gate. These simulations yielded a conclusion that to
reproduce the DC-RF dispersion the charge had to be trapped on both sides of the gate as
well as under the gate, with the majority of the charge on the drain side and that the surface
charge distributions should follow the pattern of “high electron sheet density at the gate
and low sheet density extending away from the gate over several gate lengths”. Since this
is the pattern that is also followed by an exponential function with a negative argument, we
made a hypothesis that the trapped surface charge could be following an exponential
distribution. Performing a large number of simulations with the exponential charge
7 Conclusions 143
distribution, employing the scripts developed for the purpose of automating and evaluating
a set of simulations described in Chapter 3, we eventually found the values of the fitting
parameters that reproduced the measured pulsed I-V characteristics showing current
collapse accurately. In order to reproduce the I-V characteristics of a degraded device, we
linked the electric field, via the converse piezoelectric effect, to regions of excessive
mechanical stress, under the assumptions of the „clamped model‟. Then, we attempted to
relate the calculated stress to defect density and subsequently to a charge density assumed
to be trapped in the stress generated defect states. However, this strategy was unable to
reproduce accurately the measured I-V characteristics of a degraded device. From the
insights gained in the investigation of the impact of the surface charge on the I-V
characteristics, it is obvious that, to reproduce the characteristics of the degraded device,
charge has to be trapped further away from the gate, yet the region of high stress is located
very close to the gate, which explains the failure to reproduce the measurements.
In Chapter 5 we found a surface charge distribution that reproduced the pulsed I-V
characteristics that shows current collapse. In Chapter 6, using numerical simulations, we
study the process that could result in the calibrated charge distribution. The electron
emission from the gate and trap-to-trap hopping via Poole-Frenkel emission mechanism
between the surface states was introduced and studied. Poole-Frenkel emission is strongly
nonlinear with respect to the electric field. The model for the emission rate of electrons
from the gate to the surface [132] and trap-to-trap emission frequency employs the
Eyring‟s reaction rate model [133,128], which determines the propagation velocity of the
electrons leaked from the gate, as a function of the electric field, each offering two fitting
parameters. These four fitting parameters will determine the electron leakage to the surface
and the temporal evolution of the surface charge distribution. In Chapter 6, the electron
leakage is examined for a range of values of those parameters. The charge distribution was
found to be strongly dependent on the initial surface current density of the electrons
leaking from the gate and the resulting distribution resembles the exponential distribution
only above a certain threshold value for the leakage current. By default, this model could
never result in a stationary distribution.
7 Conclusions 144
The converse piezoelectric effect was taken into account during the investigation of two
phenomena in this thesis, the electric field impact on the polarization induced bound
charge in Chapter 4, and the electric field induced mechanical stress leading to defect
formation and hence to device degradation. Yet, the electro-mechanical coupling was done
using the simplifying assumptions known as the „clamped model‟ that omits the impact of
the electric field component perpendicular to the crystal c-axis and the stresses exerted by
the surrounding matter. An indication that the „clamped model‟ is insufficient to describe
the stress field in a device is that the strain/stress measured by micro-Raman spectroscopy
is ten times higher than the calculated strain/stress, although the patterns of the strain/stress
distributions agree [30]. Therefore, it would be beneficial to perform coupled electro-
mechanical simulation of a device solving the equation (4.1), without employing the
clamped model, to obtain more realistic strain/stress distribution, including the strain
parallel with the a-axis of the crystal (and stress parallel with the c-axis), the shear
strain/stress originating from the component of the electric field perpendicular to the c-axis
and from strong gradients of the strain/stress field, and the strain/stress propagation in the
device.
Another enhancement of the work in this thesis would be to calibrate the I-V
characteristics of a degraded device searching for the appropriate distribution of the
trapped charge irrespective of the underlying mechanism, i.e. irrespective of the exact
defect distribution, and/or using the stress distribution obtained as suggested in the
previous paragraph.
The simulation of the leakage of the electrons from the gate to the device performed in
Chapter 6, took only primary leakage mechanism into account, i.e., to the surface of the
device. To obtain a more realistic picture of the charge distribution, one could simulate the
leakage including the secondary leakage mechanism, i.e. to the bulk of the device
[125,128].
Finally, another improvement of the work done in this thesis would be further automation
of the simulations. E.g. currently, the procedure is to perform all simulations and evaluate
7 Conclusions 145
all simulations thereafter. A more efficient method would of course be to perform one or
several simulations, evaluate the results, modify the parameters accordingly and perform
the simulations until a goal, i.e. a calibration of some sort, is achieved. This could be done
using either a gradient method, which is easier to code, but has the disadvantage that it may
get stuck in a local extreme, or using a genetic algorithm, which is more demanding on
computational time, but scans the phase space of parameters more efficiently.
References
[2] U. K. Mishra, Shen L., T. E. Kazior, and Y.-F. Wu, "GaN-Based RF Power Devices and Amplifiers,"
Proceedings of the IEEE, vol. 96, no. 2, pp. 287-305, 2008.
[3] E. O. Johnson, "Physical limitation on frequency and power parameters of transistors," RCA Review,
vol. 26, pp. 163-177, 1965.
[7] G. Sasso, N. Rinaldi, G. Matz, and C. Jungemann, "Analytical Models of Effective DOS, Saturation
Velocity and High-Field Mobility for SiGe HBTs Numerical Simulation," in International Conference
on Simulation of Semiconductor Processes and Devices, SISPAD, 2010, pp. 15-B.2.
[9] A. Bykhovski, B. Gelmont, and M. Shur, "The influence of the strain-induced electric field on the
charge distribution in GaN-AlN-GaN structure," Journal of Applied Physics, vol. 74, no. 11, pp. 6734-
6739, 1993.
References 147
[10] M. Asif Khan, J. N. Kuznia, J. M. Van, N. Pan, and J. Carter, "Observation of a two-dimensional
electron gas in low pressure metalorganic chemical vapor deposited GaN-AlxGa1-xN heterojunctions,"
Applied Physics Letters, vol. 60, no. 24, pp. 3027-3029, 1992.
[11] M. Asif Khan, A. Bhattarai, J. N. Kuznia, and D. T. Olson, "High electron mobility transistor based on
a GaN-AlxGa1-xN heterojunction," Applied Physics Letters, vol. 63, no. 9, pp. 1214-1215, 1993.
[12] Y.-F. Wu, B. P. Keller, S. Keller, D. Kapolnek, S. P. Denbaars, and U. K. Mishra, "Measured
microwave power performance of AlGaN/GaN MODFET," Electron Device Letters, IEEE, vol. 17, no.
9, pp. 455-457, 1996.
[13] U. K. Mishra, P. Parikh, and Y.-F. Wu, "AlGaN/GaN HEMTs-an overview of device operation and
applications," Proceedings of the IEEE, vol. 90, no. 6, pp. 1022-1031, 2002.
[15] S. Nakamura, "III-V nitride based light-emitting devices," Solid State Communications, vol. 102, no. 2-
3, pp. 237-243, 1997, Highlights in Condensed Matter Physics and Materials Science.
[16] P. Sandvik, K. Mi, F. Shahedipour, R. McClintock, A. Yasan, P. Kung, and M. Razeghi, "AlxGa1-xN
for solar-blind UV detectors," Journal of Crystal Growth, vol. 231, no. 3, pp. 366-370, 2001.
[18] M. J. Rosker, "The Present State of the Art of Wide-Bandgap Semiconductors and Their Future," in
IEEE Radio Frequency Integrated Circuits (RFIC) Symposium, 2007, pp. 159-162.
[20] J. A. del Alamo and J. Joh, "GaN HEMT reliability," Microelectronics Reliability, vol. 49, pp. 1200-
1206, 2009.
[22] A. Chini, M. Esposto, G. Meneghesso, and E. Zanoni, "Evaluation of GaN HEMT degradation by
means of pulsed I-V, leakage and DLTS measurements," Electronics Letters, vol. 45, no. 8, pp. 426-
427, 2009.
[23] M. Faqir, G. Verzellesi, A. Chini, F. Fantini, F. Danesin, G. Meneghesso, E. Zanoni, and C. Dua,
"Mechanisms of RF Current Collapse in AlGaN-GaN High Electron Mobility Transistors," Device and
Materials Reliability, IEEE Transactions on, vol. 8, no. 2, pp. 240-247, 2008.
[24] G. Meneghesso, F. Rampazzo, P. Kordoš, G. Verzellesi, and E. Zanoni, "Current Collapse and High-
Electric-Field Reliability of Unpassivated GaN/AlGaN/GaN HEMTs," Electron Devices, IEEE
Transactions on, vol. 53, no. 12, pp. 2932-2941, 2006.
[25] G. Verzellesi, R. Pierobon, F. Rampazzo, G. Meneghesso, A. Chini, U.K. Mishra, C. Canali, and E.
Zanoni, "Experimental/numerical investigation on current collapse in AlGaN/GaN HEMT's," in IEDM.
International Electron Devices Meeting, 2002, pp. 689-692.
[26] G. L. Bilbro and R. J. Trew, "RF knee walkout and source access region of unpassivated HFETs,"
Electronics Letters, vol. 42, no. 24, pp. 1425-1426, 2006.
[27] A. Sarua, Hangfeng Ji, M. Kuball, M. J. Uren, T. Martin, K. P. Hilton, and R. S. Balmer, "Integrated
micro-Raman/infrared thermography probe for monitoring of self-heating in AlGaN/GaN transistor
structures," Electron Devices, IEEE Transactions on, vol. 53, no. 10, pp. 2438-2447, 2006.
[29] J. Joh and J. A. del Alamo, "Mechanisms for Electrical Degradation of GaN High-Electron Mobility
Transistors," in IEDM. International Electron Devices Meeting, 2006, pp. 1-4.
[30] A. Sarua, Hangfeng Ji, M. Kuball, M. J. Uren, T. Martin, K. J. Nash, K. P. Hilton, and R. S. Balmer,
"Piezoelectric strain in AlGaN/GaN heterostructure field-effect transistors under bias," Applied Physics
Letters, vol. 88, no. 10, p. 103502, 2006.
[31] J. Joh and J. A. del Alamo, "Critical Voltage for Electrical Degradation of GaN High-Electron
Mobility Transistors," Electron Device Letters, IEEE, vol. 29, no. 4, pp. 287-289, 2008.
[32] E. Arslan, S. Bütün, and E. Ozbay, "Leakage current by Frenkel-Poole emission in Ni/Au Schottky
contacts on Al0.83In0.17N/AlN/GaN heterostructures," Applied Physics Letters, vol. 94, no. 14, p.
142106, 2009.
References 149
[33] O. Mitrofanov and M. Manfra, "Dynamics of trapped charge in GaN/AlGaN/GaN high electron
mobility transistors grown by plasma-assisted molecular beam epitaxy," Applied Physics Letters, vol.
84, no. 3, pp. 422-424, 2004.
[34] A. M. Wells, M. J. Uren, R. S. Balmer, K. P. Hilton, T. Martin, and M. Missous, "Direct demonstration
of the 'virtual gate' mechanism for current collapse in AlGaN/GaN HFETs," Solid-State Electronics,
vol. 49, no. 2, pp. 279-282, 2005.
[35] O. Mitrofanov and M. Manfra, "Poole-Frenkel electron emission from the traps in AlGaN/GaN
transistors," Journal of Applied Physics, vol. 95, no. 11, pp. 6414-6419, 2004.
[36] J. Frenkel, "On Pre-Breakdown Phenomena in Insulators and Electronic Semi-Conductors," Physical
Review, vol. 54, no. 8, pp. 647-648, 1938.
[37] M. Kočan, "AlGaN/GaN MBE 2DEG Heterostructures: Interplay between Surface-, Interface- and
Device- Properties," Rheinisch-Westfälische Technische Hochschule, Aachen, PhD Thesis 2003.
[38] O. Ambacher, "Growth and applications of Group III-nitrides," Journal of Physics D: Applied Physics,
vol. 31, no. 20, p. 2653, 1998.
[40] R. M. Martin, "Piezoelectricity," Physical Review B, vol. 5, no. 4, pp. 1607-1613, 1972.
[41] E. S. Hellman, "The Polarity of GaN: A Critical Review," MRS. Internet Journal of Nitride
Semiconductor Research, vol. 3, no. 11, pp. 1-11, 1998.
[42] Y. Dora, "Understanding material and process limits for high breakdown voltage AlGaN/GaN
HEMTs," University of California, Santa Barbara, PhD Thesis 2006.
[46] J. F. Nye, Physical Properties of Crystals: Their Representation by Tensors and Matrices.: Oxford
University Press, USA, 1985.
[47] F. Bernardini, "Spontaneous and Piezoelectric Polarization: Basic Theory vs. Practical Recipes," in
Nitride Semiconductor Devices: Principles and Simulation, J. Piprek, Ed.: Wiley-VCH Verlag GmbH
\& Co. KGaA, 2007, ch. 3, pp. 49-68.
[49] F. Bernardini and V. Fiorentini, "First-principles calculation of the piezoelectric tensor d of III-V
nitrides," Applied Physics Letters, vol. 80, pp. 4145-4147, 2002.
[50] S. Muensit, E. M. Goldys, and I. L. Guy, "Shear piezoelectric coefficients of gallium nitride and
aluminum nitride," Applied Physics Letters, vol. 75, no. 25, pp. 3965-3967, 1999.
[51] A. F. Wright, "Elastic properties of zinc-blende and wurtzite AlN, GaN, and InN," Journal of Applied
Physics, vol. 82, no. 6, pp. 2833-2839, 1997.
[52] A. Polian, M. Grimsditch, and I. Grzegory, "Elastic constants of gallium nitride," Journal of Applied
Physics, vol. 79, no. 6, pp. 3343-3344, 1996.
[53] V. Fiorentini, F. Bernardini, and O. Ambacher, "Evidence for nonlinear macroscopic polarization in
III-V nitride alloy heterostructures," Applied Physics Letters, vol. 80, no. 7, pp. 1204-1206, 2002.
[55] F. Bernardini and V. Fiorentini, "Nonlinear Behavior of Spontaneous and Piezoelectric Polarization in
III-V Nitride Alloys," Physica Status Solidi (a), vol. 190, no. 1, pp. 65-73, 2002.
[56] T. Palacios and U. K. Mishra, "AlGaN/GaN High Electron Mobility Transistors," in Nitride
Semiconductor Devices: Principles and Simulation, J. Piprek, Ed.: Wiley-VCH Verlag GmbH \& Co.
KGaA, 2007, ch. 10, pp. 211-233.
References 151
[57] A. Rizzi, R. Lantier, F. Monti, H. Lüth, F. Della Sala, A. Di Carlo, and Paolo Lugli, "AlN and GaN
epitaxial heterojunctions on 6H-SiC(0001): Valence band offsets and polarization fields," Journal of
Vacuum Science and Technology B, vol. 17, no. 4, pp. 1674-1681, 1999.
[58] A. Satta, V. Fiorentini, A. Bosin, Meloni F., and D. Vanderbilt, "Gallium Nitride and Related
Compounds," in Materials Research Society Symposium Proceedings, vol. 395, 1996, p. 515.
[59] F. Bernardini and V. Fiorentini, "Macroscopic polarization and band offsets at nitride heterojunctions,"
Physical Review B, vol. 57, no. 16, pp. R9427-R9430, 1998.
[60] F. Bernardini, V. Fiorentini, and D. Vanderbilt, "Spontaneous polarization and piezoelectric constants
of III-V nitrides," Physical Review B, vol. 56, no. 16, pp. R10024-R10027, 1997.
[61] F. Bernardini and V. Fiorentini, "Nonlinear macroscopic polarization in III-V nitride alloys," Physical
Review B, vol. 64, no. 8, p. 085207, 2001.
[65] W. R. L. Lambrecht, K. Kim, S. N. Rashkeev, and B. Segall, "Electronic and optical properties of the
group III-nitrides, their heterostructures and alloys," in Materials Research Society Symposium
Proceedings, vol. 395, 1996, p. 455.
transistors," Applied Physics Letters, vol. 77, no. 2, pp. 250-252, 2000.
[68] R. Vetury, N.Q. Zhang, S. Keller, and U.K. Mishra, "The impact of surface states on the DC and RF
characteristics of AlGaN/GaN HFETs," Electron Devices, IEEE Transactions on, vol. 48, no. 3, pp.
560-566, 2001.
[69] J. M. Tirado, J. L. Sanchez-Rojas, and J. I. Izpura, "Simulation of surface state effects in the transient
response of AlGaN/GaN HEMT and GaN MESFET devices," Semiconductor Science and Technology,
vol. 21, no. 8, pp. 1150-1159, 2006.
[70] G. Meneghesso, G. Verzellesi, R. Pierobon, F. Rampazzo, A. Chini, U.K. Mishra, C. Canali, and E.
Zanoni, "Surface-related drain current dispersion effects in AlGaN-GaN HEMTs," Electron Devices,
IEEE Transactions on, vol. 51, no. 10, pp. 1554-1561, 2004.
[71] G. Koley, V. Tilak, L. F. Eastman, and M. G. Spencer, "Slow transients observed in AlGaN/GaN
HFETs: effects of SiNx passivation and UV illumination," Electron Devices, IEEE Transactions on,
vol. 50, no. 4, pp. 886-893, 2003.
[72] J. M. Tirado, J. L. Sanchez-Rojas, and J. I. Izpura, "Trapping Effects in the Transient Response of
AlGaN/GaN HEMT Devices," IEEE Transactions on Electron Devices, vol. 54, pp. 410-417, 2007.
[75] J. W. Chung, J.-K. Lee, E. L. Piner, and T. Palacios, "Seamless On-Wafer Integration of Si(100)
MOSFETs and GaN HEMTs," Electron Device Letters, IEEE, vol. 30, no. 10, pp. 1015-1017, 2009.
[76] B. M. Green, K. K. Chu, E. M. Chumbes, J. A. Smart, J. R. Shealy, and L. F. Eastman, "The effect of
surface passivation on the microwave characteristics of undoped AlGaN/GaN HEMTs," Electron
Device Letters, IEEE, vol. 21, no. 6, pp. 268-270, 2000.
[77] Y. Ando, Y. Okamoto, H. Miyamoto, T. Nakayama, T. Inoue, and M. Kuzuhara, "10-W/mm AlGaN-
GaN HFET with a field modulating plate," Electron Device Letters, IEEE, vol. 24, no. 5, pp. 289-291,
2003.
[78] A. Chini, D. Buttari, R. Coffie, S. Heikman, S. Keller, and U. K. Mishra, "12 W/mm power density
References 153
AlGaN/GaN HEMTs on sapphire substrate," Electronics Letters, vol. 40, no. 1, pp. 73-74, 2004.
[79] Y.-F. Wu, M. Moore, A. Saxler, T. Wisleder, and P. Parikh, "40-W/mm Double Field-plated GaN
HEMTs," in 64th Device Research Conference, 2006, pp. 151-152.
[80] Y.-F. Wu, A. Saxler, M. Moore, R. P. Smith, S. Sheppard, P. M. Chavarkar, T. Wisleder, U. K. Mishra,
and P. Parikh, "30-W/mm GaN HEMTs by field plate optimization," Electron Device Letters, IEEE,
vol. 25, no. 3, pp. 117-119, 2004.
[81] Z. Chen, Y. Pei, S. Newman, R. Chu, D. Brown, R. Chung, S. Keller, S. P. Denbaars, S. Nakamura,
and U. K. Mishra, "Growth of AlGaN/GaN heterojunction field effect transistors on semi-insulating
GaN using an AlGaN interlayer," Applied Physics Letters, vol. 94, no. 11, p. 112108, 2009.
[83] Y.-F. Wu, D. Kapolnek, J. P. Ibbetson, P. Parikh, B. P. Keller, and U. K. Mishra, "Very-high power
density AlGaN/GaN HEMTs," Electron Devices, IEEE Transactions on, vol. 48, no. 3, pp. 586-590,
2001.
[84] V. Kumar, G. Chen, S. Guo, B. Peres, I. Eliasevich, and I. Adesida, "Field-plated 0.25 mu;m gate-
length AlGaN/GaN HEMTs on 6H-SiC with power density of 9.1 W/mm at 18 GHz," Electronics
Letters, vol. 41, no. 19, pp. 1080-1081, 2005.
[88] D. C. Dumka, C. Lee, H. Q. Tserng, P. Saunier, and M. Kumar, "AlGaN/GaN HEMTs on Si substrate
with 7 W/mm output power density at 10 GHz," Electronics Letters, vol. 40, no. 16, pp. 1023-1024,
2004.
References 154
[90] V. Kumar, A. Kuliev, R. Schwindt, M. Muir, G. Simin, J. Yang, M. Asif Khan, and I. Adesida, "High
performance 0.25 [mu]m gate-length AlGaN/GaN HEMTs on sapphire with power density of over 4.5
W/mm at 20 GHz," Solid-State Electronics, vol. 47, no. 9, pp. 1577-1580, 2003.
[91] C. Lee, P. Saunier, J. Yang, and M. A. Khan, "AlGaN-GaN HEMTs on SiC with CW power
performance of >4 W/mm and 23% PAE at 35 GHz," Electron Device Letters, IEEE, vol. 24, no. 10,
pp. 616-618, 2003.
[94] R. Aubry, J.-C. Jacquet, J. Weaver, O. Durand, P. Dobson, G. Mills, M.-A. di Forte-Poisson, S.
Cassette, and S.-L. Delage, "SThM Temperature Mapping and Nonlinear Thermal Resistance
Evolution With Bias on AlGaN/GaN HEMT Devices," Electron Devices, IEEE Transactions on, vol.
54, no. 3, pp. 385-390, 2007.
[95] B. Benbakhti, A. Soltani, K. Kalna, M. Rousseau, and J.-C. De Jaeger, "Effects of Self-Heating on
Performance Degradation in AlGaN/GaN-Based Devices," Electron Devices, IEEE Transactions on,
vol. 56, no. 10, pp. 2178-2185, 2009.
[96] S. C. Binari, P. B. Klein, and T. E. Kazior, "Trapping effects in GaN and SiC microwave FETs,"
Proceedings of the IEEE, vol. 90, no. 6, pp. 1048-1058, 2002.
[98] C. P. Wen, "Proposed GaN HFET current collapse mechanism," in Microwave Conference
Proceedings, 2005. APMC 2005. Asia-Pacific Conference Proceedings, vol. 3, 2005, pp. 4-7.
[101] Synopsys Inc., "Sentaurus Workbench User Guide, Version A-2007.12," Fremont, California,
December 2007.
[102] Synopsys Inc., "Sentaurus Structure Editor User Guide, Version A-2007.12," Fremont, California,
December 2007.
[103] Synopsys Inc., "Sentaurus Device User Guide, Version A-2007.12," Fremont, California, December
2007.
[104] S. Selberherr, Analysis and Simulation of Semiconductor Devices.: Springer-Verlag Wien, New York,
1984.
[106] D. M. Caughey and R. E. Thomas, "Carrier mobilities in silicon empirically related to doping and
field," Proceedings of the IEEE, vol. 55, no. 12, pp. 2192-2193, 1967.
[107] D. Qiao, L. S. Yu, S. S. Lau, J. M. Redwing, J. Y. Lin, and H. X. Jiang, "Dependence of Ni/AlGaN
Schottky barrier height on Al mole fraction," Journal of Applied Physics, vol. 87, no. 2, pp. 801-804,
2000.
[108] E. T. Yu, X. Z. Dang, L. S. Yu, D. Qiao, P. M. Asbeck, S. S. Lau, G. J. Sullivan, K. S. Boutros, and J.
M. Redwing, "Schottky barrier engineering in III-V nitrides via the piezoelectric effect," Applied
Physics Letters, vol. 73, no. 13, pp. 1880-1882, 1998.
[109] A. F. M. Anwar and E. W. Faraclas, "Schottky barrier height in GaN/AlGaN heterostructures," Solid-
State Electronics, vol. 50, pp. 1041-1045, 2006.
[112] R. Gaska, J. W. Yang, A. Osinsky, Q. Chen, M. Asif Khan, A. O. Orlov, G. L. Snider, and M. S. Shur,
"Electron transport in AlGaN-GaN heterostructures grown on 6H-SiC substrates," Applied Physics
Letters, vol. 72, no. 6, pp. 707-709, 1998.
[113] U. V. Bhapkar and M. S. Shur, "Monte Carlo calculation of velocity-field characteristics of wurtzite
GaN," Journal of Applied Physics, vol. 82, no. 4, pp. 1649-1655, 1997.
[117] C. H. Oxley and M. J. Uren, "Measurement of unity gain cutoff frequency and saturation velocity of a
GaN HEMT transistor," Electron Devices, IEEE Transactions on, vol. 52, no. 2, pp. 165-170, 2005.
[120] J. Zelenka, Piezoelectric resonators and their applications.: Elsevier Science Publishers B. V., 1986.
[121] B. Jogai, J. D. Albrecht, and E. Pan, "Effect of electromechanical coupling on the strain in AlGaN/GaN
heterojunction field effect transistors," Journal of Applied Physics, vol. 94, no. 6, pp. 3984-3989, 2003.
[122] T. B. Bahder, "Converse piezoelectric effect in [111] strained-layer heterostructures," Physical Review
B, vol. 51, no. 16, pp. 10892-10896, 1995.
[123] D. Balaz, K. Kalna, M. Kuball, M. J. Uren, and A. Asenov, "Impact of the field induced polarization
space-charge on the characteristics of AlGaN/GaN HEMT: Self-consistent simulation study," in IWN.
International Workshop on Nitrides, 2008.
[124] D. Balaz, K. Kalna, M. Kuball, M. J. Uren, and A. Asenov, "Impact of the field induced polarization
space-charge on the characteristics of AlGaN/GaN HEMT: Self-consistent simulation study," Physica
References 157
[125] R. J. Trew, Y. Liu, W. Kuang, and G. L. Bilbro, "Reliability modeling of high voltage AlGaN/GaN and
GaAs field-effect transistors," in Proc. of SPIE, vol. 6894, 2008, p. 68941H.
[126] D. Balaz, K. Kalna, M. Kuball, M. J. Uren, and A. Asenov, "Systematic simulation study of the impact
of virtual gate geometry on the current collapse in AlGaN/GaN HEMTs," in UK Semiconductors,
2009.
[127] W. S. Tan, M. J. Uren, P. A. Houston, R. T. Green, R. S. Balmer, and T. Martin, "Surface leakage
currents in SiNx passivated AlGaN/GaN HFETs," Electron Device Letters, IEEE, vol. 27, no. 1, pp. 1-
3, 2006.
[128] W. Kuang, R. J. Trew, and G. L. Bilbro, "An Analytical Model For Surface Leakage Currents of
AlGaN/GaN HFETs and Effects Upon Device Reliability," in WOCSDICE. Workshop On Compound
Semiconductor Devices and Integrated Circuits in Europe, 2008, pp. 133-134.
[129] D. Balaz, K. Kalna, M. Kuball, D. G. Hayes, M. J. Uren, and A. Asenov, "Impact of surface charge on
the I-V characteristics of an AlGaN/GaN HEMT," in WOCSDICE. Workshop On Compound
Semiconductor Devices and Integrated Circuits in Europe, 2009.
[131] A. Koudymov, M. S. Shur, and G. Simin, "Compact Model of Current Collapse in Heterostructure
Field-Effect Transistors," Electron Device Letters, IEEE, vol. 28, no. 5, pp. 332-335, 2007.
[132] H. Zhang, E. J. Miller, and E. T. Yu, "Analysis of leakage current mechanisms in Schottky contacts to
GaN and Al0.25Ga0.75N/GaN grown by molecular-beam epitaxy," Journal of Applied Physics, vol. 99,
no. 2, p. 023703, 2006.
[133] Khairurrijal, F.A. Noor, and Sukirno, "Modeling of stress-induced leakage current in thin gate oxides,"
in ICSE. International Conference on Semiconductor Electronics, Penang, Malaysia, 2002, pp. 375-
377.