Chapter 3
Chapter 3
Chapter 3
1 Introduction
It was found by Zhang et al. [1] in 1990 that ultrashort electromagnetic ra-
diation with terahertz (THz) bandwidth can be generated by illuminating
semiconductor surfaces with femtosecond laser pulses. Soon, the investiga-
tions of THz radiation from various semiconductors followed with study into
the emission efficiency as well as the emission mechanism [2, 3, 4, 5, 6, 7, 8, 9].
Currently, it is commonly understood that the emission of THz radiation
from semiconductor surfaces is primarily due to the surge current normal to
the surface [1, 10] and/or the second-order nonlinear optical processes in the
semiconductors [11, 12, 13]. The radiation intensity is generally low compared
to the other THz radiation sources, such as biased photoconductive (PC)
antennas, or phase-matched nonlinear optical crystals. However, after it was
revealed that the InAs surface is an efficient THz emitter and that the ef-
ficiency is further enhanced by applying an external magnetic field [14, 15],
THz radiation from an InAs surface attracted much interest because of the
potential as a simple and powerful THz radiation source in practical applica-
tions, such as spectroscopy and imaging. The special topic on the enhance-
ment of THz radiation from InAs under a magnetic field is treated in the
Chapter by Ohtake et al.
In this Chapter, we discuss THz radiation from semiconductor surfaces by
excitation of femtosecond laser pulses with special attention on the influence
of lattice vibrations and plasma oscillations in semiconductors.
K. Sakai (Ed.): Terahertz Optoelectronics, Topics Appl. Phys. 97, 63–98 (2005)
© Springer-Verlag Berlin Heidelberg 2005
64 Ping Gu and Masahiko Tani
Fig. 1. Experimental setup for the generation of THz radiation from the semicon-
ductor surface and its detection
such as a PC antenna is used, and the optical time delay of the probe laser
pulse to the pump is scanned to obtain the THz radiation waveforms (the
details of the PC antennas and their properties are described in the Chapter
by Sakai et al.).
For the explanation of the mechanisms of THz radiation from semicon-
ductor surfaces, we describe the studies carried out by the authors on sev-
eral representative semiconductors. The sample semiconductors used in the
studies are listed in Table 1 with their basic properties and some important
experimental parameters: an n-type and p-type of InSb (100) with a carrier
concentration of n ≈ 1016 cm−3 , an n-type InAs (111) and p-InAs (100) with
a carrier concentration of n ≈ 1018 cm−3 , and an n-type and p-type of InP
(100) with a residual carrier concentration of n ≈ 1018 cm−3 .
(2)
(2)
Pi (Ω) = ε0 χijk (Ω = ω1 − ω2 )
j,k
∞
· Ej (ω1 = Ω + ω2 )Ek (ω2 ) dω2 , (2)
−∞
(2)
where χijk is the second-order nonlinear susceptibility tensor for a difference
frequency, Ω = ω1 − ω2 , and Ej (ω1 ) (Ek (ω2 )) is the amplitude spectral com-
ponent of the pump laser at frequency ω1 (ω2 ) in the j(k) direction. Here,
i, j, and k are the dummy indices for x-, y-, and z-directions in the crys-
tallographic axis system. The integral is extended to the negative frequency
by using the definition, E(−ω) = E ∗ (ω) (the asterisk means the complex
conjugate).
Terahertz Radiation from Semiconductor Surfaces 67
ob
In the far field, the observed THz field amplitude, ETHz , is proportional
to the projection of the second time derivative of the nonlinear polarization
to the polarization direction of detection, e (a unit vector normal to the
observation direction), at the observation point:
∂ 2 P (t)
ob
ETHz (t) = e · E THz ∝ e · . (3)
∂t2
ob
The spectral amplitude, ETHz (Ω), at a frequency Ω is thus given by
ob
ETHz (Ω) ∝ Ω 2 e · P (Ω) . (4)
The most unambiguous evidence for the contribution of the χ(2) process
is a strong dependence of the emitted THz radiation intensity on the crystal
orientation to the pump-laser polarization. By rotating a sample about its
surface normal, the relative contribution of the azimuthal-angle-dependent
DFM component to the total THz radiation can be estimated.
Figure 2 shows the azimuthal-angle dependence of the peak amplitude
in the time-resolved waveform of THz radiation (a) from n-InAs(111) and
(b) from n-InSb(100) surfaces measured in the experimental configuration
shown in Fig. 1 when the samples were rotated about their surface normal.
A laser with a 40 fs pulse width, a 76 MHz repetition rate, and a central
wavelength of 800 nm was focused on the sample semiconductor surfaces at
45◦ by using a lens with a focal length of about 25 cm. The pump laser
was p-polarized, and the average power was about 80 mW. The excitation
intensity on the samples were estimated to be about 60 MW/cm2 .
Taking the surface normal as the X-axis and the reflection plane as the
XY -plane in the laboratory frame, the nonlinear polarization induced in the
68 Ping Gu and Masahiko Tani
semiconductor due to the optical rectification for (111) and (100) surfaces is
expressed by (5) and (6), respectively [19]:
⎛ ⎞
Px (Ω)
P = ⎝Py (Ω)⎠
Pz (Ω)
⎛ ⎞
− √13 cos2 φ + √26 sin2 φ
⎜ ⎟
= 2ε0 d14 E 2 (Ω) ⎝ √26 cos2 φ cos 3θ − √26 cos φ sin φ⎠ , (5)
√2 cos2 φ sin 3θ
6
⎛ ⎞
Px (Ω)
P = ⎝Py (Ω)⎠
Pz (Ω)
⎛ 2 ⎞
cos φ sin 2θ
= 2ε0 d14 E 2 (Ω) ⎝ sin 2φ sin 2θ ⎠ , (6)
sin 2φ cos 2θ
∞
E (Ω) =
2
E(ω1 = Ω + ω2 )E(ω2 ) dω2
−∞
∞
= E(ω1 = Ω − ω2 )E ∗ (ω2 ) dω2 . (7)
−∞
Here, φ is the angle between the surface normal and the pump-laser beam
refracted inside the sample, θ is the azimuthal angle of the sample orientation
around the X-axis (the angle between the Y -axis in laboratory frame and the
crystallographic (11 − 2) direction), E(ωi ) is the amplitude component of the
(2)
pump laser for frequency ωi , and d14 = χ14 /2 is the nonlinear susceptibility
coefficient for the difference frequency, Ω, in the contracted notation, E(Ω)
is the autocorrelation function of E(ω).
Using (3) and considering the refraction at the interface between the semi-
conductor/air interface, the p-polarized THz field amplitude observed in the
direction of optical reflection is given by the following equation:
ob
ETHz ∝ eP = (− sin φTHz , cos φTHz , 0)(PX , PY , PZ )t
= −PX sin φTHz + PY cos φTHz , (8)
where φTHz is the refraction angle of THz radiation inside the semiconduc-
tor (see Fig. 3). The refraction angles for the optical and THz beams are
determined by the generalized Snell’s law as
where nopt and nTHz is the refractive index for the pump laser and THz ra-
diation in the semiconductor, respectively. For the pump-laser wavelength of
Terahertz Radiation from Semiconductor Surfaces 69
800 nm, φ is 10.9◦ , 9.0◦ and 11.8◦ for InAs, InSb, and InP, respectively. For
THz radiation, φTHz is estimated to be 10.9◦ , 21.5◦, and 11.9◦ for InAs, InSb,
and InP, respectively. Using these values, the azimuthal-angle dependence of
the radiation amplitude due to the optical rectification can be written as
follows:
ob
ETHz ∝ 1.093d14(cos 3θ − 0.103) for (111)-InAs , (10a)
ob
ETHz ∝ 0.1823d14 sin 2θ for (100)-InAs , (10b)
ob
ETHz ∝ −0.069d14 sin 2θ for (100)-InSb , (10c)
ob
ETHz ∝ 0.199d14 sin 2θ for (100)-InP . (10d)
In semiconductors with a wide bandgap, such as GaAs (Eg = 1.43 eV) or InP
(Eg = 1.34 eV), the surface bands of a semiconductor lie within its energy
bandgap, and thus Fermi-level pinning occurs, leading to band bending and
formation of a depletion region, where the surface built-in field exists [21].
After optical excitation, the electrons and holes are accelerated in opposite
directions under the surface-depletion field, forming a surge current in the
direction normal to the surface. The direction and magnitude of the surface-
depletion field depend on the dopant or impurity species and the position of
the surface states relative to the bulk Fermi level. In general, the energy band
is bent upward in n-type semiconductors (Fig. 4a) and downward in p-type
semiconductors (Fig. 4b).
Therefore, the built-in surface field in p-type semiconductors drives the
photogenerated carriers, and thus the transient surge current, in the opposite
direction to that in n-type semiconductors as shown in Fig. 4a. In the far-
field approximation, the emitted THz-radiation-field amplitude, ETHz (t), is
proportional to the time derivative of the surge current, J(t):
∂J(t)
ETHz (t) ∝ . (11)
∂t
When the depletion-surface field is the dominant mechanism for the surge
current, the polarity of the emitted THz radiation waveform is opposite be-
tween that of the n-type and that of the p-type semiconductors.
On the other hand, when the photo-Dember effect is the dominant mech-
anism for the surge current, the polarity of the THz radiation will remain the
Terahertz Radiation from Semiconductor Surfaces 71
Fig. 4. Band diagram and the schematic flow of drift current in (a) n-type and
(b) p-type semiconductors
InAs and InSb are very interesting semiconductors because of their high elec-
tron mobilities: ≈ 30 000 cm2/(V · s) for InAs and ≈ 76 000 cm2/(V · s) for
InSb. Recently, InAs is attracting much attention as a THz emitter since
it has been found that the THz emission power from InAs is significantly
enhanced under magnetic fields [15]. The surface-depletion voltage of the
narrow-bandgap semiconductors is generally not so large because of the small
bandgap energy. For the excitation of narrow-bandgap semiconductors with
NIR light (hν ≈ 1.5 eV), the absorption depth is very small (≈ 100 nm) [22],
and the excess energy of photoexcited carriers is very large. All these con-
ditions in the narrow-bandgap semiconductors enhance the photo-Dember
72 Ping Gu and Masahiko Tani
kB (Te b − Th ) 1 (b + 1)∆n
VD = ln 1 + . (13)
e b+1 n0 b + p 0
Here, b = µe /µp is the mobility ratio of the electrons (µe ) and holes (µp ),
n0 and p0 are the initial density of the electrons and holes, Te and Th are
the temperature of photoexcited electrons and holes, respectively. This equa-
tion tells us that the photo-Dember effect is enhanced by larger electron
mobility (µe ∝ b), and higher electron excess energy (∝ Te ). The narrow-
bandgap semiconductors have the preferred conditions necessary to create
a large photo-Dember field, that is, the very large electron mobilities and
large excess carrier energies. Moreover, the photo-Dember field (VD /d, d : ab-
sorption depth) in narrow-bandgap semiconductors is further enhanced by
the small absorption depth. On the other hand, the surface-depletion field
is expected to be small because of the small bandgap energy, in contrast to
the widegap semiconductors. For 800 nm light, the absorption depth for InSb
and InAs is estimated to be 94 nm and 142 nm, respectively, while that of
InP is 304 nm. For 800 nm light excitation (hν = 1.55 eV), The excess en-
ergy, ∆E, in InSb and InAs is 1.38 eV and 1.18 eV, respectively, which are
much bigger values than that of InP (∆E = 0.21 eV), because of their nar-
row bandgaps (see Table 1). These conditions indicate the existence of a large
photo-Dember field for InSb and InAs under 800 nm optical excitation. Thus,
the main source of THz radiation for InAs and InSb is considered to be the
74 Ping Gu and Masahiko Tani
Fig. 7. Time-domain waveforms of the THz radiation (a) from n- and p-InSb, and
(b) n- and p-InAs. The azimuthal angle of crystal orientation was 26◦ for n-InAs
(111), and 90◦ for the other (100)-oriented samples to suppress the contribution
from the optical rectification effect
photo-Dember field, rather than the screening of the surface depletion field.
The radiation amplitude of the n-type InAs and n-type InSb is a little bigger
than their counterpart p-type samples. This suggests that there might be an
enhancement of THz radiation due to the surface-depletion field because the
direction of the surface depletion field and the photo-Dember field is the same
in n-type semiconductors.
From (13), we can deduce several important properties for the emission
of THz radiation by the photo-Dember effect.
1. Low residual electron and hole concentrations (n0 and p0 ) enhance the
photo-Dember effect.
2. The photo-Dember voltage, VD , and thus the emitted THz-field ampli-
tude, is expected to be proportional to the pump intensity, I, (VD ∝
∆n ∝ I) in a low-intensity regime and proportional to ln(I) [∝ ln(∆n)]
in a high-intensity regime.
Terahertz Radiation from Semiconductor Surfaces 75
Fig. 9. Time-domain
waveforms of n-InAs
(111) excited by 780 nm
and 1.55 µm light
Some readers might be uncomfortable with the above explanation of THz ra-
diation from the longitudinal polarization due to LO phonons since LO
phonons, in general, do not interact with electromagnetic waves, which are
essentially transverse waves. This conceptual conflict is reconciled by inter-
preting the coherent LO phonon described above as the upper branch of the
phonon–polariton mode at low wave numbers, q 0; When the wave number
of optical phonons, q, is exactly zero, we can not distinguish the transverse
mode from the longitudinal mode (and vice versa), and thus the two modes
are degenerated, having the same frequency, ωLO , as the frequency of unper-
turbed LO phonons [28]. In the excitation conditions with a finite incident
angle, the excited longitudinal polarization gets a small but finite wavevector
component in the transverse direction due to the phase delay determined by
78 Ping Gu and Masahiko Tani
(a)
TO-phonon
div E = ρP/ε0 = 0 k
ionic
E displacement
ular to the c-axis. The samples of PbTe and CdTe had surfaces perpen-
dicular to the 100 direction. Te has several optical-phonon modes be-
cause of the low symmetry of its crystal structure (hexagonal): a Raman-
active A1 mode (3.6 THz), two Raman-and-IR-active E modes (ETO/LO :
2.76/3.09 THz, ETO/LO : 4.22/4.26 THz), and one only-IR-active A2 mode
(A2 ,TO/LO: 2.6/2.82 THz). PbTe has the NaCl-like crystal structure and
has a TO- and an LO-phonon mode at 0.96 THz and 3.42 THz, respectively.
CdTe has the zincblende crystal structure and has TO- and an LO-phonon
modes at 4.20 THz and 5.08 THz, respectively. [35]
The absorption of the THz radiation through the thin emitting volume is
too small to explain the spectral dip observed at TO-phonon frequencies: the
absorption of radiation at TO-phonon frequencies by propagation through a
distance equal to the optical absorption depth is estimated to be < 1% for
Te and 1% to 2% for PbTe. For CdTe, the absorption of the radiation at the
TO-phonon frequency is estimated to be about 10% for a propagation length
of 150 nm, which is half the thickness of the depletion layer calculated using
the Schottky-barrier model. Thus, we need to find other mechanisms for the
spectral dip at TO-phonon frequencies.
θi ,
where t(p) is the transmission coefficient for the p-polarized wave, θd is the
refracted angle given by (14), and Ei is the far field of the dipole 2 in air.
With an incident angle θa the Fresnel transmission coefficient t(p) is given by
2 cos θa
t(p) (θa ) = . (16)
ñd cos θa + cos θd
Using these equations, we find that the radiation power Pa (θa ), emitted by
dipole 1 to the air at an angle θa and per unit solid angle, as
cε0 E12 cε0 Ei2
Pa (θa ) = = |t(p) (θa )|2 sin2 θd
2 2
3 sin2 θd cos2 θa 3
= P0 ≡ P0 ga (θa ) , (17)
8π |ñd cos θa + cos θd | 2 8π
where c is the velocity of light and P0 = 4πcε0 Ei2 /3 is the total power emit-
ted by a dipole when it is placed in free space. Here, ga (θa ) represents the
Terahertz Radiation from Semiconductor Surfaces 87
ga (θa ) cos2 θa
= =
ga (θa ) + gd (θd ) cos2 θa + n3d |ñd |2 cos2 θd
1
= , (19)
1 + n3d |ñd |2 cos2 θd /cos2 θa
where the angular integrations, air dΩ and diel dΩ, are taken for the air-
side half-space and dielectric-side half-space, respectively, and the relation
between θa and θd is given by (14). To obtain the second equation, we as-
sumed the array factors, fa (θ, φ) and fd (θ, φ), are delta functions centered
at the direction to the optical reflection and refraction, δ(θa − θ)δ(φ) and
δ(θd − θ)δ(φ), respectively. To a rough approximation, (19) reduces to
−5/2
Ga (ω) ∼ ñ−5
d = εd . (20)
88 Ping Gu and Masahiko Tani
−3/2
This relation corresponds to Ga ∝ ε−sd = εd given for a dipole antenna on
a dielectric substrate oriented parallel to the dielectric surface. The reason
why the power coefficient, s, is larger for the vertical dipole is attributed to
the fact that the vertical dipole emits more power in the direction parallel to
the surface and thus is more subject to the total reflection at the air/dielectric
interface. Equation (20) suggests that the THz-emission efficiency from semi-
conductor surfaces is strongly influenced by the dispersion property of the
semiconductor.
For the frequency-dependent refractive index , ñd (ω), we assume
(εs − ε∞ )ωTO
2
ñ2d (ω) = εd (ω) = ε∞ + , (21)
ωTO − ω − iωΓ
2 2
Fig. 14. Frequency of the LO-phonon–plasmon mode in GaAs plotted against the
carrier density: ωLO = 8.76 THz, ωTO = 8.06 THz, ε∞ = 10.6
where, N is the total density of electrons including both the intrinsic electrons
and photoexcited electrons, and m∗ is the reduced mass of electron.
In polar materials, LO-phonons and -plasmons couple and give rise to the
LO-phonon–plasmon modes, whose frequencies are dependent on the carrier
density N . In Fig. 14, the oscillation frequencies of the coupled modes without
damping are shown as a function of the square-root of plasma density, N 1/2 ,
for GaAs. The frequency of TO-phonons and LO-phonons in GaAs is ωTO =
8.06 THz and ωLO = 8.76 THz, respectively. When the carrier density is low
and the frequency of the plasmon oscillations is far below the LO-phonon
frequency, the interaction between plasmons and LO-phonons is weak. In this
case, the upper branch of the LO-phonon–plasmon coupled modes (L+ mode
with frequency ω = ω+ ) has phonon-like characteristics with its frequency
close to the pure LO-phonon (ω+ ∼ ωLO ), and the lower branch of the coupled
modes (L− mode with frequency ω = ω− ) has plasmon-like characteristics
with its frequency close to the pure plasmon (ω− ∼ ωpl ). As the carrier
density increases, as does the plasmon frequency, L− , L+ modes anticross
near the LO-phonon frequency due to the LO-phonon–plasmon interaction. √
At higher carrier density the L+ mode frequency increases linearly with N ,
while the L− mode frequency approaches that of the TO-phonon (ω− ∼ ωTO ).
In this case, the L+ mode has plasmon characteristics with its frequency close
to the plasma frequency (ω+ ∼ ωpl ), while the L− mode is considered as the
longitudinal phonon mode, whose frequency is softened due to the screening
90 Ping Gu and Masahiko Tani
Fig. 15. (a) Differential transient reflectivity for n-GaAs (N > 1018 cm−3 ) and
(b) its Fourier-transformed spectrum
the high density of the modes at these frequencies, agreeing with the exper-
imental observation. Figure 15a shows the differential photoreflectance of a
heavily doped n-GaAs (N > 1018 cm−3 ). In the Fourier-transformed spec-
trum (Fig. 15b) we observe a dominant spectral peak at the TO-phonon
frequency, which is interpreted as the L− mode of the LO-phonon–plasmon
coupled mode. The same discussion as for GaAs is valid for other polar semi-
conductors. Therefore, we expect to observe the THz emission originating
from the L− mode with the TO-phonon frequency at high carrier densities
(N > 1018 cm−3 ). Such a coherent emission near the TO-phonon frequency
is suppressed due to the high refractive index and is thus difficult to observe.
However, for moderately doped polar semiconductors the L− or the L+ mode
frequencies are offresonant from the TO-phonon frequency and thus are ob-
served as coherent THz emission with femtosecond laser excitation.
Gu et al. [46] carried out THz-emission spectroscopy for doped InSb films
and observed coherent THz radiation corresponding to L− and/or the L+
phonon–plasmon modes. In the experiment they used a 1.4 µm thick InSb
film with an intrinsic carrier concentration of n = (6.70 ×1016 )cm−3 fab-
ricated on GaAs (100) substrates by MOCVD. Raman-scattering measure-
ments showed that the TO- and LO-phonon modes of the film were at frequen-
cies of 5.39 THz and 5.72 THz, respectively. The electron plasma frequency
at zero wave number (q = 0) is calculated to be 4.45 THz, which is slightly
lower than the LO-phonon frequency of InSb. From the intrinsic carrier con-
centration the L− and L+ modes are calculated to be 3.94 THz and 6.10 THz
at q = 0. To detect high-frequency THz radiation, ultrashort laser pulses less
than 30 fs and tight focusing of the gate laser beam on a 5 µm LTG-GaAs
PC gap of the detector antenna were used [19].
Figure 16 shows the THz waveforms detected by a PC antenna in the
(a) reflection (THz radiation in the optical-reflection direction is detected)
and (b) transmission geometry (THz radiation transmitted through the sub-
strate is detected) for the same InSb film at an optically excited carrier den-
sity nex ≈ (5 × 1016 ) cm−3 . The inset in Fig. 16a is the waveform magnified
by a factor of 3 and shows the oscillatory structure after a 3.5 ps delay time.
As shown in the inset, a beating is observed, indicating the presence of two
modes with different frequencies. The corresponding Fourier-transformed am-
plitude spectra of the THz radiation are shown in Figs. 17a and b. The first
large and broad spectral peak at around 1 THz corresponds to the first cycle
of the THz pulse originating from the surge current induced by the photo-
Dember effect. To remove the contribution from the THz radiation due to the
surge current, which forms the slowly varying background in the spectrum,
we carried out a Fourier transformation with a time window later than the
third maximum (≈ 3.5 ps) in the waveform in Figs. 16a and b. The spec-
tral peak profiles by the windowed Fourier transformations are also shown
92 Ping Gu and Masahiko Tani
Fig. 18. Spectra measured (a) in the reflection geometry and (b) transmission
geometry at optical pump powers of 60 mW, 40 mW and 20 mW. To remove the
contribution of the THz-radiation component due to the surface surge current, the
Fourier transformations of waveforms were carried out with limited time windows.
The frequencies of the L+ mode and L− mode calculated for the intrinsic carrier
density are indicated by the vertical arrows
4 Conclusions
The THz radiation from semiconductor surfaces is attractive not only as
a simple THz-radiation source but also as a THz-radiation source for high
frequencies (> 3 THz), where other radiation sources such as PC antennas
are not efficient.
Not only as a simple THz-radiation source, THz radiation from semi-
conductor surfaces can be used as the probe for the investigation of carrier-
transport dynamics near to semiconductor surfaces [50, 51, 52]: Since the
Terahertz Radiation from Semiconductor Surfaces 95
References
[1] X.-C. Zhang, B. Hu, J. Darrow, D. Auston: Appl. Phys. Lett. 56, 1011 (1990)
63
[2] X.-C. Zhang, J. Darrow, B. Hu, D. Auston, M. Schmidt, P. Tham, E. Yang:
Appl. Phys. Lett. 56, 2228 (1990) 63
[3] X.-C. Zhang, B. Hu, S. Xin, D. Auston: Appl. Phys. Lett. 57, 753 (1990) 63
[4] X.-C. Zhang, D. Auston: J. Appl. Phys. 71, 326 (1992) 63
[5] X.-C. Zhang, Y. Jin, K. Yang, L. Schowalter: Phys. Rev. Lett. 69, 2303 (1992)
63
[6] M. Li, F. Sun, G. Wagoner, M. Alexander, X.-C. Zhang: Appl. Phys. Lett. 67,
25 (1995) 63
[7] J. Perdersen, I. Balslev, J. Hvam, S. Keiding: Appl. Phys. Lett. 61, 1372 (1992)
63
[8] B. Hu, X.-C. Zhang, D. Auston: Appl. Phys. Lett. 57, 2629 (1990) 63
[9] Y. Jin, X. Ma, G. Wagoner, M. Alexander, X.-C. Zhang: Appl. Phys. Lett. 65,
682 (1994) 63
[10] T. Dekorsy, H. Auer, C. Waschke, H. Bakker, H. Roskos, H. Kurz, V. Wanger,
P. Grosse: Phys. Rev. Lett. 74, 738 (1995) 63, 80, 81
[11] S. Chuang, S. Schmitt-Rink, B. Greene, P. Saeta, A. Levi: Phys. Rev. Lett.
68, 102 (1995) 63, 66
96 Ping Gu and Masahiko Tani
[33] S. Ruhman, A. Joly, K. Nelson: IEEE J. Quantum Electron. 24, 460 (1988)
79
[34] M. Tani, R. Fukasawa, H. Abe, S. Matsuura, K. Sakai: J. Appl. Phys. 83, 2473
(1998) 80
[35] P. Grosse, W. Richter: Numerical Data and Functional Relationships in Sci-
ence and Technology, Landoldt-Börnstein, New Series 17 (Springer, Berlin,
Heidelberg 1983) 81
[36] H. Zeiger, J. Vidal, T. Cheng, E. Ippen, G. Dresselhaus, M. Dresselhaus: Phys.
Rev. B 45, 768 (1992) 83
[37] T. Pfeifer, T. Dekorsy, W. Kütt, H. Kurz: Appl. Phys. A 55, 482 (1992) 84
[38] A. Kuznetsov, C. Stanton: Phys. Rev. B 51, 7555 (1995) 84, 90, 95
[39] G. Cho, T. Dekorsy, H. Bakker, R. Hovel, H. Kurz: Phys. Rev. Lett. 77, 4062
(1996) 84
[40] M. Hase, K. Mizoguchi, H. Harima, F. Miyamaru, S. Nakashima, R. Fukasawa,
M. Tani, K. Sakai: J. Lumin. 76/77, 68 (1998) 84
[41] D. Rutledge, D. Neikirk, D. Kasilingam: Infrared Millimeter Waves 10, 1
(1983) 85, 86
[42] W. Stutzman, G. Thiele: Antenna Theory and Design (Wiley, New York 1981)
Chap. 3 87
[43] S. Perkowitz: Phys. Rev. B 12, 3210 (1975) 88
[44] S. Perkowitz, R. Thorland: Phys. Rev. B 9, 545 (1974) 88
[45] P. Grosse, W. Richter: Phys. Stat. Sol. 41, 239 (1970) 88
[46] P. Gu, M. Tani, K. Sakai, T.-R. Yang: Appl. Phys. Lett. 77, 1798 (2000) 91
[47] R. Kersting, K. Unterrainer, G. Strasser, H. Kauffmann, E. Gornik: Phys. Rev.
Lett. 79, 3038 (1997) 93
[48] For review paper, see for example, G.C. Cho, T. Dekorsy, H.J. Bakker,
R. Hävel, and H. Kurz: Phys. Rev. Lett. 77, 4062 (1996), T.T. Dekorsy,
H. Auer, C. Waschke, H.J. Bakker, H.G. Roskos, and H. Kurz: Phys. Rev.
Lett. 74, 738 (1995), M. Vobebörger, H.G. Roskos, F. Wolter, C. Waschke,
and H. Kurz: J. Opt. Soc. Am. B 13, 1045 (1996) 94
[49] C. Collins, P. Yu: Solid State Commun. 51, 123 (1984) 94
[50] A. Leitenstorfer, S. Hunsche, J. Shah, M. Nuss, W. Knox: Phys. Rev. B 61,
16642 (2000) 94, 95
[51] J.-H. Son, T. Norris, J. Whitaker: J. Opt. Soc. Am. B 11, 2519 (1994) 94
[52] Y. Shimada, K. Hirakawa, S.-W. Lee: Appl. Phys. Lett. 81, 1642 (2002) ;
N. Tsukada: Jpn. J. Appl. Phys. 33, 4807 (1994), Appl. Opt. 36, 7853–7859
(1997); J. Appl. Phys. 80, 4214–4216 (1996); Opt. Lett. 15, 323–325 (1990,
1995), Opt. Lett. 62, 1265 (1993) 94
Index