Longitudinal optical (LO) phonons in GaN generated in the channel of high electron mobility transistors (HEMT) are shown to undergo nearly elastic scattering via collisions with hot electrons. The net result of these collisions is the diffusion of LO phonons in the Brillouin zone causing reduction of phonon and electron temperatures. This previously unexplored diffusion mechanism explicates how an increase in electron density causes reduction of the apparent lifetime of LO phonons, obtained from the time resolved Raman studies and microwave noise measurements, while the actual decay rate of the LO phonons remains unaffected by the carrier density. Therefore, the saturation velocity in GaN HEMT steadily declines with increased carrier density, in a qualitative agreement with experimental results.

III-Nitride based high electron mobility transistors (HEMTs) can be microwave devices of choice when both high frequency and high power operation are required.1–3 At high fields both current density and cut-off frequency of HEMTs are determined to a large degree by the saturation velocity of electrons vsat. GaN has a high breakdown field and a parabolic conduction band (CB) with satellite valleys far removed from the CB edge. Furthermore, the energy of longitudinal optical (LO) phonons is very high, ωLO=90meV. The combination of these fortuitous characteristics causes one to expect the saturation velocity in GaN to be on the scale of vsatωLO/mc=2.8×107cm/s. In reality vsat is less than that, especially when the electron density ne is high, i.e., in the regime where the power-amplifying HEMT is expected to operate.1–3 Our most recent experiments4 have shown that vsat steadily falls from about 1.9 to less than 0.9×107cm/s as ne is increased from 1018cm2 to 1019cm2 as shown in Fig. 1. Here ne is the “average” 3D density, estimated from the 2D density n2D also shown in the figure and the full width at half maximum of the wavefunction using Schrodinger-Poisson simulation. The source of this reduction has been discussed extensively, and, while details may differ, there is a broad consensus that the reduction in saturation is caused by the hot phonons accumulating in the channel.5–7 The accumulation is caused by the fact that the lifetime of LO phonons in GaN, as measured in Raman experiments, can be as long as τLO=23ps (Refs. 8 and 9) at low ne. The lifetime is so long because ωLO in GaN is high, and an LO phonon cannot split into two acoustic (TA or LA) ones.10 The only channel available for LO decay is the slow Ridley channel,11 in which an LO phonon splits into a TO and a TA (LA) phonon. Since the group velocity of LO phonons is very small, they accumulate in the HEMT channel, causing an increase in the electron temperature Te. In addition, the momentum relaxation rate increases and the velocity saturates.5,6 Naturally, the higher ne, the higher the density of LO phonons, which in turn reduces the vsat.

FIG. 1.

Charge density (ne) dependent electron saturation velocity (vsat) extracted from measured velocity-field characteristics in AlGaN/GaN HEMT in Ref. 4. Here, ne is the “average” 3D density estimated from the 2D density n2D (also shown in the figure) and the full width at half maximum of the wavefunction using Schrodinger-Poisson simulation. The error bars include the vsat values extracted from multiple devices.

FIG. 1.

Charge density (ne) dependent electron saturation velocity (vsat) extracted from measured velocity-field characteristics in AlGaN/GaN HEMT in Ref. 4. Here, ne is the “average” 3D density estimated from the 2D density n2D (also shown in the figure) and the full width at half maximum of the wavefunction using Schrodinger-Poisson simulation. The error bars include the vsat values extracted from multiple devices.

Close modal

This simple explanation assumes that τLO is constant and does not depend on electron concentration and temperature. It is understood that electrons are not directly involved in the decay of LO phonons. This commonsense assumption of constant τLO has been challenged in the last decade. The time resolved Raman measurements12 in which decay times of the Anti-Stokes line, caused by hot phonons and thus associated with τLO, has been shown to decrease from 2.5 ps to less than 300 fs as ne increased from 1017 to 1019 cm−3. Apparent decrease of τLO was also inferred from the measurements of Te,via hot-electron noise at microwave frequencies.13–16 Since hot electron and hot phonons are at thermal equilibrium with each other, the rate of hot electrons cooling had been associated with LO phonon lifetime. This measured lifetime had also shown a decrease with increased ne and Te.

The mechanism of the decrease in τLO with increase of ne is not obvious,17 since at first look only the anharmonicity of lattice vibrations defines τLO. In 2006 Dyson and Ridley18 came up with an ingenious explanation based on coupling between LO phonons and plasmons. The resonant frequency of plasmons near the Brillion zone (BZ) center is ωp=(e2ne/εmc)1/2, where mc=0.2m0 is the CB effective mass, and ε=8.9 is a low frequency dielectric constant. When ne approaches 1019 cm−3, ωp becomes comparable with ωLO, and LO phonon and plasmon couple forming two hybridized phonon-plasmon (HPP) longitudinal modes with frequencies ω+ and ω. Higher energy HPP ω+ has large longitudinal electric field and thus efficiently scatters electrons. Unlike LO phonons, the HPP mode is dispersive and has a group velocity on the scale of 106 cm/s. There are two ostensive reasons why the HPP's decay faster than “bare” LO phonons.18–20 First, the rate of Ridley process of decay, LO → TO + LA is proportional to LA frequency, ωA1=ωLOωTO, and to the LA phonon density of states, which according to Debye model is ρA1ωA12. Then, the LO phonon lifetime is τLO(ωLOωTO)3, and when the LO phonon hybridizes with a plasmon, the lifetime of the HPP τ+(ω+ωTO)3 decreases dramatically in comparison with that of a bare LO phonon simply because ω+>ωLO. The second reason for apparent decrease of LO phonon lifetime is the migration of coupled plasmon-phonon quasiparticles out of the HEMT channel. Indeed, with the group velocity of 106 cm/s it would take only 300 fs for the hot quasiparticles to move out of the channel.

However, both of these explications appear questionable when subjected to in-depth examination. First of all, when one considers real phonon dispersion in GaN,21 at frequencies near and above ωA1=ωLOωT0=240cm1, the acoustic phonon density no longer follows a square law, and, in fact, decreases with frequency as evidenced in Fig. 2(a). Therefore, when hybridization takes place, the density ρA2 of the acoustic modes with frequency ωA2=ω+ωT0>240cm1 into which HPP decays is actually reduced compared with ρA1; therefore, no decrease in the lifetime can take place. When it comes to the “migration” explanation, it is important to note that HPP frequency ω+(x) is coordinate dependent and differs from ωLO only where ne is large, i.e., the inside of the channel as shown in Figs. 1(b) and 1(c). Therefore, the HPP moving ballistically towards the edge of the channel simply gets reflected back into the channel, as shown in Fig. 1(c), as propagating further would clearly violate energy conservation. Alternatively, one can say, that an HPP mode whose dispersion is parabolic (i.e., electron-like) simply forms a confined state inside the channel, described by the wave function ψ+(x) and similar in all respects to the confined electron state. The HPP's in this state cannot possibly exit the channel of HEMT in the vertical direction. Furthermore, if τLO goes down as drastically as predicted by the “plasmon hybridization” model, then vsat should not change with ne changing from 1018 to 1019 cm−3.18 Yet our data in Fig. 1 (Ref. 4) does show significant decrease of vsat in that density range confirming earlier results,22 albeit in disagreement with results in Refs. 23 and 24 where the electron density was determined in a different way. Finally, apparent decrease of optical phonon lifetime with ne had also been measured in silicon,25 and germanium,26 non-polar materials where plasmons and phonons do not couple at all.

FIG. 2.

(a) Decay of bare LO phonon ωLO (solid arrows) or hybrid phonon-plasmon (HPP) mode ω+ (dashed arrows) into a TO phonon ωTO and acoustic phonon ωA1 or ωA2 with densities of states ρA1 or ρA2. Reprinted with permission from V. Yu. Davydov, Yu. E. Kitaev, I. N. Goncharuk, A. N. Smirnov, J. Graul, O. Semchinova, D. Uffmann, M. B. Smirnov, A. P. Mirgorodsky, and R. A. Evarestov, Phys. Rev. B 58, 12899 (1998). Copyright 1998 American Physical Society. (b) CB diagram and electron density ne in the HEMT channel. (c) Diagram of the HPP energy ω+(x) in the HEMT channel and the wavefunction of the HPP mode ψ+(x) formed by the reflection on the channel boundaries.

FIG. 2.

(a) Decay of bare LO phonon ωLO (solid arrows) or hybrid phonon-plasmon (HPP) mode ω+ (dashed arrows) into a TO phonon ωTO and acoustic phonon ωA1 or ωA2 with densities of states ρA1 or ρA2. Reprinted with permission from V. Yu. Davydov, Yu. E. Kitaev, I. N. Goncharuk, A. N. Smirnov, J. Graul, O. Semchinova, D. Uffmann, M. B. Smirnov, A. P. Mirgorodsky, and R. A. Evarestov, Phys. Rev. B 58, 12899 (1998). Copyright 1998 American Physical Society. (b) CB diagram and electron density ne in the HEMT channel. (c) Diagram of the HPP energy ω+(x) in the HEMT channel and the wavefunction of the HPP mode ψ+(x) formed by the reflection on the channel boundaries.

Close modal

To reconcile these facts, it is essential to recognize that characteristics measured in Raman and noise experiments are different from actual phonon lifetimes. In Raman experiments, only the dynamics of phonons with wavevector q ∼ 0 is observed, while noise temperature measurements determine only the maximum LO phonon population near q0=2mcωLO/0.68nm1, where most of LO phonons are produced. But electron-phonon interaction occurs throughout a much wider span of wave vectors in the BZ. Hence, if there existed an ne-dependent mechanism instigating migration or “diffusion” of the LO phonons throughout the BZ, this diffusion would quickly deplete the LO phonon population in the region where they are produced (q0 in Raman experiments and qq0 in electron noise measurements). Measured, or apparent, LO phonon lifetime τapp then would be much lower than τLO. With the total number of hot LO phonons in the BZ steadily growing as ne increases, vsat would continue its steady decrease, albeit the rate of this decrease would experience some slow down, as measured in our experiments.

In this work we establish that this mechanism causing the decrease of τapp with ne does exist and is nothing but the scattering of LO phonons on hot electrons, a second-order process that can be remarkably efficient due to large density of LO phonon states. The “classical” diagram of the process is shown in Fig. 3(a) where an LO phonon with quasi-momentum q1 and hot electron k1 “collide” and the new states carry momenta q2=q1+Δq and k2=k1Δq. Since the energy of LO phonon barely changes, the process is nearly elastic, in fact similar to Rayleigh scattering of photons in the atmosphere. The quantum nature of the process is shown in Fig. 3(b), where the distribution of hot electrons in the CB is shown as a tinted sphere, with the initial k1 and final k2 states located on the same equal energy surface. According to 2-nd order perturbation theory, scattering involves absorption of the initial phonon q1 and electron making a transition into the virtual state k=k1+q1 followed by emission of phonon q2 and electron going into the final state k2, or the emission of q2 occurs first as electron makes transition to a different virtual state, k=k1q2. According to Fermi golden rule27 and diagram in Fig. 3(b), the total rate of scattering between two LO phonon states q1 and q2 is

(1)

where V is the volume, 1/ε=1/ε1/ε, and ε is the dielectric constant in the visible/infrared region. The two terms in the parenthesis correspond to two different quantum paths, and their interference cancels divergences at q1,2=0. Since δ(Ek2Ek1)=mc/(2k1Δq)δ(cosθΔq/2k1), only the electron states with energies higher than Emin(Δq)=2(Δq)2/8mc can scatter the LO phonons by Δq; therefore, scattering exhibits a strong dependence on Te. As explained earlier, for large ne the LO phonon hybridizes with a plasmon and one should use the HPP mode frequency ω+ in place of ωLO, which should slightly enhance scattering, but we neglect it as well as degeneracy of the electron gas at ne exceeding 1019 cm−3 in our order-of-magnitude analysis. To proceed further it is convenient to introduce normalized wave vector qr=q/q0, normalized temperature Tr=kBTe/ωLO, and normalized electron density, Nr=neq03. Then, performing the summation in (1), which involves averaging over the angles, one obtains the estimate of the rate of the net scattering for the phonons from the state with a given value of q

(2)

where nq is the number of LO phonons in the state q, Gq is the generation rate, which includes the net LO phonon creation by the electrons as well as their possible generation via Stokes Raman scattering, n¯=[exp(ωLO/kBT)1]1 is the LO equilibrium number at the lattice temperature T, and R0=π1/2(e2q0/2πεωLO)2ωLO = 0.66×1015s1. In other words, if a single LO phonon had been generated, then for electron concentration of say, 1019cm3 (Nr0.03) and Te of 1000 K (Tr1) the apparent lifetime of it would have been only about 50 fs or less. Of course, as LO phonons are generated throughout the BZ, the scattering out of a given state q will be counter-acted by scattering into this state, so the net rate will be less and the process will acquire the features of diffusion. Indeed, if one assumes that the phonons are distributed in the BZ according to Gaussian law, nq=n0exp(qr2/2σq2), where σq2=q2nq=0, for relatively small q2nq the integral in (2) can be evaluated as (3/2σq2)Tr2 and then one obtains dnqdtDq2nqnqn¯τLO where the diffusion coefficient in reciprocal space is Dq=(3/2)R0Tr3/2Nr, or Dq=(3/2)R0(kBT/ωLO)3/2neq01 in absolute units of cm−2 s−1.

FIG. 3.

(a) “Classical” representation of electron-phonon elastic collision (b) quantum picture of LO-scattering on hot electrons involving intermediate virtual electron states k′ and k″.

FIG. 3.

(a) “Classical” representation of electron-phonon elastic collision (b) quantum picture of LO-scattering on hot electrons involving intermediate virtual electron states k′ and k″.

Close modal

To illustrate how scattering by hot electrons determines the apparent lifetime of LO phonons, we have solved (2) for the case where the LO phonons are generated near the center of the BZ as nq=n0exp(qr2/2σq2) where σq=1 (or σq=q0 in absolute units), Te=600K, ne=5×1017cm3, and τLO=2.5ps. The first order processes of phonon creation and annihilation, contained in the term Gq, can only maintain this distribution and not alter it, and of course they cannot create the LO phonons at large q-vectors. Therefore, in our calculations we only consider the actual decay of phonons via elastic scattering by electrons and via normal TO + LA (TA) process. In Fig. 4(a), the snapshots of non-equilibrium LO phonon distribution nqn¯ from t = 0 to t = 2 ps are shown. The LO phonon density near the center of the BZ decays exponentially with τapp=1.95ps. In Fig. 4(b), we plot the density of per unit interval of q, i.e., (nqn¯)qr2, and one can clearly see that before LO phonons decay they diffuse away from the center of the BZ. Since τapp1=τLO1+τdiff1, the diffusion time is τdiff9ps, i.e., diffusion is not significant. However, when ne is increased to 1019cm3, the apparent LO phonon decay accelerates dramatically as shown in Figs. 3(c) and 3(d). Near the BZ center τapp decreases to about 0.25 ps. This is the lifetime observed in Raman experiments.12 In the vicinity of q0 one can see τapp(q0)0.5ps. This is what is observed in the noise measurement.13–15 The diffusion time near the BZ center is τdiff0.29ps indicating that diffusion plays the dominant role. From Figs. 3(b) and 3(d) one can also see that the total number of excess LO phonons, the area under the curve, decays with the constant lifetime τLO independent of ne.

FIG. 4.

Temporal evolution of non-equilibrium LO phonon distribution for an electron temperature of 600 K and electron densities of (a) and (b) ne=5×1017cm3 and (c) and (d) ne=1×1019cm3.

FIG. 4.

Temporal evolution of non-equilibrium LO phonon distribution for an electron temperature of 600 K and electron densities of (a) and (b) ne=5×1017cm3 and (c) and (d) ne=1×1019cm3.

Close modal

Fig. 5 demonstrates electron-density-dependence of τapp near the BZ center for different Te's. More than a tenfold decrease of τapp occurs between ne=1018cm3 and ne=1019cm3, which is in line with the lifetime decreases reported in Refs. 12–16. The dependence of τapp on Te is strong at high density, when τapp is dominated by diffusion, although it is somewhat weaker than Te3/2 dependence predicted using simplified equations earlier. Also plotted are the data from Raman measurements in Ref. 12. Agreement is good; however, one must not overlook that in addition to obvious electron concentration and temperature dependencies, the diffusion time is also strongly influenced by how the generated LO phonons are distributed in BZ, for which we have chosen a Gaussian distribution with σq=1. Under different excitation conditions, the observed τapp may vary from one experiment to another. In fact, the only way to accurately predict τapp would be to obtain a self-consistent solution for dynamics of both hot LO phonons and hot electrons in each point in the BZ. Such a task is beyond the scope of this paper. Nevertheless, we can say, with confidence that the main experimentally observed trend, the order-of-magnitude reduction of τapp for ne exceeding 1018 cm−3, can be faithfully replicated using realistic input parameters σq and Te. Showing that diffusion in reciprocal space can lead to tenfold decrease in τapp is an important step towards a full understanding of hot phonon dynamics.

FIG. 5.

Calculated apparent lifetime of LO phonon near the BZ center as a function of electron density for three different temperatures. Black squares—experimental data from Ref. 12.

FIG. 5.

Calculated apparent lifetime of LO phonon near the BZ center as a function of electron density for three different temperatures. Black squares—experimental data from Ref. 12.

Close modal

At first, the finding that LO phonon scattering by electrons occurs at rates that are only a few times slower than the rates of LO phonon generation in HEMT (typically tens of femtoseconds5,6) can be surprising. Scattering is a 2nd-order process, while generation is a 1st order one. When dealing with, say photons, the higher order processes are usually orders-or-magnitudes weaker, but phonons have exceptionally large density of states; therefore, different order processes may have comparable strength. For instance, four–phonon scattering processes can play nearly as important a role as three phonon processes in reducing thermal conductivity.27 

Besides being a new interesting physical phenomenon, phonon scattering by hot electrons plays a key role in determining vsat. The total number of LO phonons in the BZ is unaffected by scattering. Therefore as ne increases vsat steadily declines, in agreement with the experimental data.4,22 But re-distribution of LO phonons reduces their maximum number in the vicinity of q0, which in turn reduces Te (as evidenced from noise measurements) and makes cooling more efficient. In addition, the LO phonons that spread further towards the BZ edges do not interact with electrons as strongly as the ones near q0, so the momentum scattering rate decreases.28 As a result the absolute value of the slope of vsat vs. ne is gradually reduced, although the slope always remains negative in full agreement with Fig. 1.

In conclusion, we have shown that LO phonon scattering by hot electrons in GaN causes diffusion of LO phonons in the BZ. This is manifested in the reduction of the measured apparent lifetime of LO phonons with an increase in electron density. Phonon diffusion strongly affects electron temperature and velocity saturation. Once it is fully incorporated into a self-consisted model of HEMTs, it will help to improve the design of the future III-Nitride HEMTs and other high speed devices, as nothing discussed here is limited to nitrides.

The authors gratefully acknowledge the support from DATE MURI (ONR N00014-11-1-0721, Dr. Paul Maki).

1.
Y.
Yue
,
Z.
Hu
,
J.
Guo
,
B.
Sensale-Rodriguez
,
G.
Li
,
R.
Wang
,
F.
Faria
,
T.
Fang
,
B.
Song
,
X.
Gao
,
S.
Guo
,
T.
Kosel
,
G.
Snider
,
P.
Fay
,
D.
Jena
, and
H.
Xing
,
IEEE Electron Device Lett.
33
(
7
),
988
990
(
2012
).
2.
K.
Shinohara
,
D.
Regan
,
A.
Corrion
,
D.
Brown
,
Y.
Tang
,
J.
Wong
,
G.
Candia
,
A.
Schmitz
,
H.
Fung
,
S.
Kim
, and
M.
Micovic
,
IEEE Int. Electron Devices Meet.
2014
,
27.2.1
27.2.4
.
3.
Y.
Tang
,
K.
Shinohara
,
D.
Regan
,
A.
Corrion
,
D.
Brown
,
J.
Wong
,
A.
Schmitz
,
H.
Fung
,
S.
Kim
, and
M.
Micovic
,
IEEE Electron Device Lett.
36
(
6
),
549
551
(
2015
).
4.
S.
Bajaj
,
O. F.
Shoron
,
P. S.
Park
,
S.
Krishnamoorthy
,
F.
Akyol
,
T.-H.
Hung
,
Sh.
Reza
,
E.
Chumbes
,
J.
Khurgin
, and
S.
Rajan
,
Appl. Phys. Lett.
107
,
153504
(
2015
).
5.
B. K.
Ridley
,
W. J.
Schaff
, and
L. F.
Eastman
,
J. Appl. Phys.
96
(
3
),
1499
1502
(
2004
).
6.
J.
Khurgin
,
Y.
Ding
, and
D.
Jena
,
Appl. Phys. Lett.
91
(
25
),
252104
(
2007
).
7.
J. H.
Leach
,
C. Y.
Zhu
,
M.
Wu
,
X.
Ni
,
X.
Li
,
J.
Xie
,
Ü.
Özgür
,
H.
Morkoç
,
J.
Liberis
,
E.
Šermukšnis
,
A.
Matulionis
,
H.
Cheng
, and
Ç.
Kurdak
,
Appl. Phys. Lett.
95
,
223504
(
2009
).
8.
K. T.
Tsen
,
D. K.
Ferry
,
A.
Botchkarev
,
A.
Serdlov
,
A.
Salvador
, and
H.
Morkoc
,
Appl. Phys. Lett.
72
,
2132
(
1998
).
9.
L.
Shi
,
F. A.
Ponce
, and
J.
Menendez
,
Appl. Phys. Lett.
84
,
3471
(
2004
).
10.
P. G.
Klemens
,
Phys. Rev.
148
,
845
(
1966
).
11.
B. K.
Ridley
,
J. Phys.: Condens. Matter
8
,
L511
(
1996
).
12.
K. T.
Tsen
,
J. G.
Kiang
,
D. K.
Ferry
, and
H.
Morkoc
,
Appl. Phys. Lett.
89
,
112111
(
2006
).
13.
A.
Matulionis
,
J.
Liberis
,
I.
Matulionienė
,
M.
Ramonas
,
L. F.
Eastman
,
J. R.
Shealy
,
V.
Tilak
, and
A.
Vertiatchikh
,
Phys. Rev. B
68
,
035338
(
2003
).
14.
A.
Matulionis
,
J.
Liberis
,
E.
Šermukšnis
,
J.
Xie
,
J. H.
Leach
,
M.
Wu
, and
H.
Morkoç
,
Semicond. Sci. Technol.
23
,
075048
(
2008
).
15.
A.
Matulionis
,
J. Phys.: Condens. Matter
21
,
174203
(
2009
).
16.
A.
Matulionis
,
J.
Liberis
,
I.
Matulioniene
,
M.
Ramonas
,
E.
Šermukšnis
,
J. H.
Leach
,
M.
Wu
,
X.
Ni
,
X.
Li
, and
H.
Morkoç
,
Appl. Phys. Lett.
95
,
192102
(
2009
).
17.
T.
Beechem
and
S.
Graham
,
J. Appl. Phys.
103
,
093507
(
2008
).
18.
A.
Dyson
and
B. K.
Ridley
,
J. Appl. Phys.
103
,
114507
(
2008
).
19.
A.
Dyson
and
B. K.
Ridley
,
J. Appl. Phys.
108
,
104504
(
2010
).
20.
A.
Dyson
and
B. K.
Ridley
,
J. Appl. Phys.
112
,
063707
(
2012
).
21.
V. Yu.
Davydov
,
Yu. E.
Kitaev
,
I. N.
Goncharuk
,
A. N.
Smirnov
,
J.
Graul
,
O.
Semchinova
,
D.
Uffmann
,
M. B.
Smirnov
,
A. P.
Mirgorodsky
, and
R. A.
Evarestov
,
Phys. Rev. B
58
,
12899
(
1998
).
22.
C. H.
Oxley
,
M. J.
Uren
,
A.
Coates
, and
D. G.
Hayes
,
IEEE Trans. Electron Devices
53
,
565
(
2006
).
23.
C. H.
Leach
,
C. Y.
Zhu
,
M.
Wu
,
X.
Ni
,
X.
Li
,
J.
Xie
,
Ü.
Özgür
,
H.
Morkoç
,
J.
Liberis
,
E.
Šermukšnis
,
A.
Matulionis
,
T.
Paskova
,
E.
Preble
, and
K. R.
Evans
,
Appl. Phys. Lett.
96
,
133505
(
2010
).
24.
A.
Matulionis
,
Semicond. Sci. Technol.
28
,
074007
(
2013
).
25.
J. J.
Letcher
,
K.
Kang
,
D. G.
Cahill
, and
D. D.
Dlott
,
Appl. Phys. Lett.
90
,
252104
(
2007
).
26.
M. L.
Ledgerwood
and
H. M.
van Driel
,
Phys. Rev. B
54
,
4926
(
1996
).
27.
J.
Ziman
,
Electrons and Phonons
(
Oxford
,
New York
,
2001
).
28.
J. B.
Khurgin
,
D.
Jena
, and
Y. J.
Ding
,
Appl. Phys. Lett.
93
,
032110
(
2008
).