Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Chapter 3

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Terahertz Radiation

from Semiconductor Surfaces

Ping Gu1,2 and Masahiko Tani1,3


1
National Institute of Information and Communications Technology
558-2 Iwaoka, Nishi-ku, Kobe, Hyogo 651-2492, Japan
2
Present address: Faculty of Education, Wakayama University
930 Sakaedani, Wakayama-shi, Wakayama 640-8510, Japan
guping@center.wakayama-u.ac.jp
3
Present address: Institute of Laser Engineering, Osaka University
2-6 Yamadaoka, Suita, Osaka 565-0871, Japan
tani@ile.osaka-u.ac.jp

Abstract. In this chapter, terahertz radiation from semiconductor surfaces by


excitation of femtosecond laser pulses is discussed with special attention, paid to
the influence of lattice vibrations and plasma oscillations in semiconductors. The
effect of dielectric dispersion near the optical-phonon resonance, the coupling of
the THz electrical transient with longitudinal optical phonons and semiconductor
plasmons is considered.

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

In Sect. 2, we describe the several basic mechanisms of THz radiation


from semiconductor surfaces. Three mechanisms are specified and explained:
the optical rectification effect, the surface depletion field, and the photo-
Dember effect. A detailed study is presented for THz radiation from InAs
and InSb, for which the contributions from the optical rectification and the
photo-Dember effect are both significant.
In Sect. 3, we describe THz radiation from coherent LO-phonons and LO-
phonon–plasmon coupled modes in polar semiconductors. The mechanism of
emission of THz radiation from coherent phonons is explained in relation
to the excitation mechanism of coherent phonons with specific examples of
Te, PbTe, and CdTe. The effect of the dielectric dispersion near the optical-
phonon resonance is discussed. A study of THz radiation from doped InSb
by the authors is also presented as an example of THz radiation from the
LO-phonon–plasmon coupled modes.

2 Terahertz Radiation from Semiconductor Surfaces


In this section we describe the basic mechanisms of emission of THz radi-
ation from semiconductor surfaces. Emission mechanisms of THz radiation
from bulk semiconductor surfaces are categorized into two types. One is the
nonlinear optical processes without real carrier excitation and the other is
the ultrafast surface surge-current with real carriers due to the surface built-
in field or photo-Dember effect. The THz emission by the nonlinear optical
process is explained as the optical rectification of ultrashort laser pulses,
which creates a transient polarization on the semiconductor surface. The
THz radiation is emitted in proportion to the second time derivative of the
electronic polarization due to the optical rectification (in the far-field ap-
proximation). This process is equivalently explained in the frequency domain
as the difference-frequency mixing (DFM) between spectrum components of
the femtosecond laser pulse in the semiconductors [12]. For the surge-current
model, two origins are considered: one is the acceleration of photoexcited car-
riers by the surface-depletion field, and the other is the photo-Dember effect
originating from the difference between the diffusion velocities of the electrons
and holes. Because electrons gain much greater velocity due to their higher
mobility than that of holes, electrons diffuse more rapidly from the surface
inward of the semiconductor, creating an effective surge current normal to
the surface.
A typical experimental setup for measurements of THz radiation gener-
ated from a semiconductor surface is illustrated in Fig. 1. Near infrared (NIR)
femtosecond light pulses from a mode-locked Ti: sapphire laser (λ ≈ 800 nm,
hν ≈ 1.5 eV, δt < 100 fs) excite the semiconductor surface at a finite incident
angle, such as 45◦ , and THz radiation emitted in the direction of optical re-
flection is collected and focused to a detector by using a pair of paraboloidal
mirrors. For the time-domain measurement, an ultrafast sampling detector,
Terahertz Radiation from Semiconductor Surfaces 65

Fig. 1. Experimental setup for the generation of THz radiation from the semicon-
ductor surface and its detection

Table 1. Sample properties

Sample InSb InAs GaAs InP


Bandgap Eg 0.17 eV 0.36 eV 1.43 eV 1.34 eV
Electron mobility µe 76 000 30 000 8600 4000
(cm2 /(V · s))
Hole mobility µh 3000 240 400 650
(cm2 /(V · s))
Electron mass m∗e 0.015me 0.027me 0.067me 0.073me
Hole mass m∗h 0.18me 0.33me 0.5me 0.40me
Refractive idex 4.47 3.73 3.68 3.47
(λ = 800 nm)
Refractive idex 1.93 3.78 3.61 3.52
(at 1 THz)
Absorption depth 94 nm 142 nm 749 nm 305 nm
(λ = 800 nm)
Excess energy ∆E 1.38 eV 1.18 eV 0.12 eV 0.21 eV
(T = Te + Th = 23 ∆E
κ
) (10 500 K) (9000 K) (900 K) (1600 K)
Electron temp. Te 9800 K 8300 K 790 K 1350 K
Hole temp. Th 700 K 700 K 110 K 250 K
66 Ping Gu and Masahiko Tani

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 .

2.1 Optical Rectification

In the case of high-intensity excitation, the optical rectification effect becomes


dominant [11]. The nonlinear optical process was first suggested by Chuang
et al. [11, 16, 17] as the emission mechanism of THz radiation from semicon-
ductor surfaces. They suggested an effective χ(2) process, which is actually
a χ(3) process involving the DC surface-depletion field, in addition to the
incident pump laser field. It was later found that the radiation emitted from
InP by excitation of high-intensity laser pulses of 2.0 eV light was due to the
bulk optical rectification (a χ(2) process) [12]. Many other semiconductors
without the inversion crystal symmetry, such as the zincblende-type semi-
conductors, have also been reported to emit THz radiation due to the bulk
optical rectification effect [17, 18]. The THz radiation field generated by the
optical rectification, E THz (t), is proportional to the second-order nonlinear
polarization (in near field), which is described by the following equations:
 +∞
1
P (t) = P (2) (Ω)e−iΩt dΩ , (1)
2π −∞

(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

Fig. 2. Azimuthal-angle dependence


of THz-emission amplitude from semi-
conductors with (a) n-InAs (111)
(squares) and (b) n-InSb (100) (cir-
cles) surfaces at a 45◦ incidence an-
gle of the excitation light on the sam-
ple. The solid lines are the theoretical
curves fitted to the data

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

sin 45◦ = nopt sin φ = nTHz sin φTHz , (9)

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

Fig. 3. Optical geometry for


generation and detection of
THz radiation from a semi-
conductor surface

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)

The nonlinear contribution is proportional to the azimuthal angular mod-


ulation of cos 3θ with a small DC offset for (111)-oriented crystals, and sin 2θ
for (100)-oriented crystals in the experimental configuration shown in Fig. 1.
The constants in front of d14 in (10a–d) represent the geometrical factors of
the nonlinear contribution for the 45◦ pump-incidence on the samples. Al-
though nonlinear optical rectification coefficients, d14 (Ω ≈ 0), for InSb, InAs
and InP are not known, the second-harmonic generation coefficients, d14 (2ω),
may be helpful in indicating the relative magnitude of d14 (Ω ≈ 0) among the
samples: d14 (2ω) at 1.06 µm is 520 pm/V ± 47 pm/V, 364 pm/V ± 47 pm/V,
and 167 pm/V for InSb, InAs, and InP, respectively [20]. Figure 2a shows
the azimuthal-angle dependence measured for the (111)-InAs sample. A pro-
nounced angle dependence is observed (squares), which agrees well with the
dependence described by (10a). From a theoretical curve fitting (solid line),
we estimated the ratio of the nonlinear contribution to the total radiation
amplitude to be 40%.
The azimuthal-angle dependence for the (100)-oriented n-InSb sample is
also shown in Fig. 2b by open circles, for which the angle dependence is
not so significant because of the small geometrical factor (≈ −0.069) for
the (100) surface. The ratio of the nonlinear contribution to the total ra-
70 Ping Gu and Masahiko Tani

diation amplitude is estimated to be 6% from the theoretical curve fitting


(solid line). It should be noted that for the (100)-oriented zincblende crystals
the small nonlinear contribution arises from the offnormal incidence of the
pump-laser beam (45◦ in this case), and the nonlinear electro-optic effect is
always zero for the normal incidence of the pump-laser beam. The azimuthal-
angle dependences for the other (100)-oriented samples were as weak as that
observed for the n-InSb sample. From the results of the azimuthal-angle de-
pendence, it is concluded that, for the offnormal incidence, the nonlinear
contribution to THz radiation is finite but not very significant for the (100)-
oriented zincblende-type semiconductor surfaces, while it can be very signif-
icant for the (111)-oriented ones.
The contribution to the THz emission from the surge current alone can
be investigated by choosing a proper azimuthal angle, where the nonlinear
contribution becomes null: For an azimuthal angle of 26◦ for (111)-InAs and
0◦ (or 90◦ ) for all (100)-oriented samples, the nonlinear contribution becomes
null (The θ-dependent factor vanishes).

2.2 Surge Current

2.2.1 Surface Depletion Field

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

same irrespective of the type of semiconductor, as explained in Sect. 2.2.2.


Therefore, by comparing the polarity of the THz waveforms emitted from an
n-type semiconductor and that of a p-type semiconductor, we can determine
which mechanism is dominant in the semiconductor.
The time-domain THz-radiation waveforms of n-InP(100) and p-InP(100)
are shown in Fig. 5. The azimuthal angle, θ, of the crystal orientation to the
pump-laser polarization was set near 90◦ for these (100)-oriented samples
to suppress the contribution from the nonlinear optical rectification effect.
It is clearly shown in Fig. 5 that the THz waveform flips its polarity be-
tween n-type and p-type InP. The contribution of the photo-Dember effect is
expected to be negligible because the absorption depth of an 800 nm pump
laser in InP is relatively long (≈ 0.3 µm) and the excess carrier energy is small
(≈ 0.3 eV) as indicated in Table 1, neither of which is a preferred condition to
form a strong photo-Dember field, as discussed in Sect. 2.2.2. Therefore, the
surge current due to the surface-depletion field is considered to be the domi-
nant mechanism for THz radiation from InP. The same discussion applies for
other wide-bandgap semiconductors, such as GaAs.

2.2.2 Photo-Dember Effect

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

Fig. 5. Time-domain wave-


forms of the THz radiation
from n- and p-InP. The az-
imuthal angle of crystal orien-
tation was 90◦ for the (100)-
oriented samples so that the
contribution from the optical
rectification effect was sup-
pressed

Fig. 6. Schematic flow of


diffusion current by photoex-
cited carriers near the sur-
face of a semiconductor

effect, which is known to generate current or voltage in semiconductors due


to the difference of the electron and hole diffusion velocities.
The diffusion current due to the photo-Dember effect after photoexcita-
tion near a semiconductor surface is illustrated in Fig. 6. Because the electron
mobility is always larger than that of holes, the direction of diffusion current
due to the photo-Dember effect is the same with any kind of semiconductor
and irrespective of the doping type (n or p). Therefore, the THz radiation
emitted from the surface surge current due to the photo-Dember effect will
show the same polarity for n-type and p-type semiconductors.
The diffusive currents of the electrons (Jn ) and holes (Jp ) are, respectively,
described by the following equations [23],
∂∆n
Jn ∼ −eDe , (12a)
∂x
∂∆p
Jp ∼ eDh , (12b)
∂x
Terahertz Radiation from Semiconductor Surfaces 73

where e is the electron unit charge, ∆n and ∆p the density of photocreated


electrons and holes, De and Dh diffusion coefficient of electrons and holes,
respectively. The diffusion coefficient D is defined by the Einstein relation,
D = kB T µ/e, where kB is the Boltzman constant, T is the temperature of
the corresponding carrier, and µ is the mobility of electrons or holes. The
THz radiation due to the Dember current Jdif = Jn + Jp is thus proportional
to the difference in the mobility and temperature for the electrons and holes,
and the gradient of the carrier density.
The time-domain waveforms for the (100)-oriented n-InSb and p-InSb are
shown in Fig. 7a, and the (111)-oriented n-InAs and the (100)-oriented p-InAs
in Fig. 7b. The azimuthal angle, θ, of the crystal orientation was set near to
26◦ for the n-InAs (111) sample, and 0◦ for the other (100)-oriented samples
so that the contribution from the optical rectification effect was suppressed.
Because the THz waveforms of the n-type and p-type of both samples have
the same polarity, and their polarity is the same as that observed for the
n-InP shown in Fig. 5, whose direction of surface-depletion field is expected
to be the same as that of the photo-Dember field, the dominant mechanism for
the surge current on InSb and InAs surfaces is considered to be an ultrafast
build-up and relaxation of the diffusion current due to the photo-Dember
effect.
We now extend our discussion by using the equation of the steady-state
photo-Dember voltage (VD ) [24]

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. 8. Time-domain wave-


forms of InAs (dotted curve)
and InSb (solid curve) at
800 nm excitation

3. The photo-Dember voltage does not depend strongly on the electron–hole


mobility ratio, b, when it is large enough (> 10). This is the case for InSb
(b = 95) and InAs (b = 125).
Because of the small absorption depth of InSb (92 nm), the photo-Dember
field (ED ∝ VD /d) should be almost twice that of InAs (142 nm) for 800 nm
light excitation. In addition, the electron mobility, µe , of InSb is about twice
that of InAs. Because the emitted THz-radiation-field amplitude is propor-
tional to the acceleration field and electron mobility (ETHz ∝ ∆J ∝ eµe ED ),
it is expected that the THz emission-efficiency of InSb is about four times
better in amplitude and therefore 16 times better in power than that of InAs.
However, the experimental results do not support this prediction.
Figure 8 shows the THz-radiation waveforms observed for InAs (dotted
curve) and InSb (solid curve) at 800 nm excitation. The THz power observed
from InSb was merely one-hundredth of that observed from InAs (without
a magnetic field). The unexpectedly low THz emission of InSb might be
explained by the reduction of the transient mobility due to intervalley scat-
terings of electrons to the L-valley, where the electron mobility is expected
to be extremely low. Because the energy of the L-valley from the top of the
valence-band edge (the Γ -point) is 1.03 eV in InSb, the electrons are easily
scattered into the L-valley by the 800 nm light excitation (hν = 1.55 eV). On
the other hand, as the energy of the L-valley is 1.53 eV (the energy barrier
is higher than this value) in InAs, the electrons are scarcely scattered to the
L-valley by the 800 nm light excitation.
Since the carrier dynamics such as the diffusions and intervalley scatter-
ings strongly depends on the excitation wavelength, the wavelength depen-
dence of THz radiation will provide important information on the emission
mechanism of THz radiation from semiconductor surfaces. Figure 9 shows
the waveforms of THz radiation from n-InAs (111) at 780 nm excitation with
a power of 3.3 mW (solid curve) and at 1.55 µm excitation with a power of
9.0 mW (dotted curve). By normalizing the amplitude signal with the respec-
tive pump and probe-beam power, the THz-emission efficiency at 1.55 µm
76 Ping Gu and Masahiko Tani

Fig. 9. Time-domain
waveforms of n-InAs
(111) excited by 780 nm
and 1.55 µm light

was estimated to be almost one order of magnitude lower than that at


780 nm. The large absorption depth (d = 590 nm) and smaller excess en-
ergy (∆E = 0.44 eV) at a 1.55 µm excitation would give about one-sixth
of the photo-Dember field expected at a 780 nm excitation (d = 140 nm,
∆E = 1.23 eV) if the simple theory of the photo-Dember effect is valid (note
that the photon number, thus the photoexcited carrier number, for 1.55 µm
light was doubled with the same power of 780 nm light). Thus, the photo-
Dember model agrees reasonably well with the experimental results at differ-
ent wavelengths for InAs. In contrast to the case of InAs, Howells et al. [25]
reported the emission efficiency of THz radiation from InSb increases by
about six times in amplitude (normalized by the photon numbers) at 1.9 µm
excitation compared to that at 800 nm. This observation is not consistent
with the photo-Dember model, which predicts a considerable decrease of the
THz emission efficiency at 1.9 µm compared to that at 800 nm because of the
large absorption depth for 1.9 µm light. Takahashi et al. [26] also reported the
THz radiation power from InSb increases by about several tens of times at
1560 nm excitation compared to that at 780 nm. However, these results can
be explained if we assume a significant reduction of the transient mobility
due to the L-valley scattering of electrons, as Howells et al. also suggested in
their paper.
In fact, Nuss et al. [27] reported a significant drop in mobility due to inter-
valley scattering in GaAs. Although further experimental studies are needed
to come to a definite conclusion. We believe the intervalley scattering plays
an important role in the THz-emission mechanism in semiconductors.

3 Terahertz Radiation from Coherent Phonons


and Plasmons
In the previous section, we have described three basic mechanisms for THz ra-
diation emitted from semiconductor surfaces: optical rectification, ultrafast
Terahertz Radiation from Semiconductor Surfaces 77

screening of the surface-depletion field, and ultrafast build-up of the photo-


Dember field, induced by femtosecond laser irradiation on semiconductor
surfaces.
In this section, we describe emission of THz radiation from oscillating
polarizations associated with coherent lattice vibrations and/or plasma oscil-
lations induced by femtosecond laser irradiation on semiconductor surfaces.
By excitation with a femtosecond laser pulse, an electrical transient is cre-
ated by the ultrafast screening of the surface-depletion field and/or ultrafast
building up of the photo-Dember field as discussed in Sect. 2.2.2. This electri-
cal shockwave triggers coherent lattice vibrations, which are called coherent
phonons. Such coherent lattice vibrations induce changes of macroscopic op-
tical constants, such as the polarizability or the linear dielectric susceptibility
(the dielectric constant). Therefore, the coherent motion of lattice vibrations
can be observed as the transient change of transmission or reflection at the
sample surface of a second laser pulse (the probe pulse) that is synchronized
with the first pump laser pulse with a variable time delay.
When the coherent phonons are infrared active, the lattice vibrations in-
duce oscillations of macroscopic polarization. As illustrated in Fig. 10, trans-
verse optical phonons (TO phonons) do not produce macroscopic charge den-
sity or macroscopic polarization because the integration of the polarization
induced by the coherent phonons over a finite thickness, which corresponds to
the absorption depth of the excitation laser, is averaged out to null (Fig. 10a).
On the other hand, the longitudinal optical phonons (LO phonons) directed
normal to the surface can create a macroscopic polarization, which is observed
as an oscillating charge on the surface (Fig. 10b). The macroscopic polariza-
tion induced by coherent LO phonons thus emits electromagnetic radiation
(THz radiation) at the LO-phonon frequency.

3.1 Coupling of Transverse Electromagnetic Waves


with Longitudinal Coherent Phonons and/or Plasmons

3.1.1 Coupling to LO Phonons

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

Fig. 10. (a) Charge density of TO pho-


non is effectively zero. (b) Charge den-
sity of LO phonon is not zero, and
macroscopic polarization is formed per-
pendicular to the semiconductor surface

the timing of the pump-laser incidence on the semiconductor surface. There-


fore, the coherent LO phonon described above is not a pure LO phonon but
rather considered as a hybrid mode of LO and TO phonons, whose trans-
verse components couple to the electromagnetic waves and are able to emit
electromagnetic waves. When the wave vector of the coupled polarization in
the transverse direction is small but nonzero, its frequency is close to, but
slightly higher than, the frequency of the unperturbed LO phonon (ωLO ).
Such an upshift of the frequency of THz radiation from an optical coherent
phonon was actually observed for Te by Dekorsy et al. [22].

3.1.2 Coupling to Plasma Oscillations

When a semiconductor with a finite carrier density is excited impulsively


with femtosecond lasers, coherent plasmon oscillations are initiated in a sim-
ilar manner as the coherent optical phonons. The plasma starts to oscillate
coherently after the excitation with a femtosecond laser pulse, due to surface-
field screening or build-up of a photo-Dember field in a direction normal to
the surface. Plasmon oscillations are longitudinal and thus can not couple to
Terahertz Radiation from Semiconductor Surfaces 79

the transverse oscillations of an electromagnetic field or can not emit electro-


magnetic waves in a homogeneous or symmetrical excitation geometry. With
a finite incident angle of excitation of a semiconductor surface, however, the
plasma oscillations can emit electromagnetic waves in the same way as dis-
cussed for the THz radiation from coherent LO phonons. The frequency of
unperturbed plasma oscillations (plasma frequency) is dependent on the car-
rier density, N , and is given by ωp = e2 N/εε0 m∗ (−e: electron charge,
ε: dielectric permeability of semiconductor, ε0 : dielectric constant of vac-
uum, and m∗ : effective electron mass). When the plasma frequency is close
to the LO-phonon frequency, the two vibrations are coupled and give rise
to coupled-mode vibrations, that is, the LO-phonon–plasmon modes. These
coupled modes are also observed as coherent emission of radiation from semi-
conductor surfaces after femtosecond laser irradiation.
THz radiation is observed not only from bulk semiconductor surfaces
but also from semiconductor surfaces with quantum-well structures. Such
THz radiation originates from charge oscillations due to two coupled quan-
tized states, or Bloch oscillations in superlattice structures. Readers who are
interested in THz radiation from semiconductor quantum-well structures are
recommended to refer to a recent review on coherent THz radiation in semi-
conductors by Gornik and Kersting [29], which also gives a good review of
THz radiation from bulk semiconductors.

3.1.3 Coherent-Phonon Excitation Mechanisms

For the later discussion, it is worth mentioning the excitation mechanisms


for coherent phonons observed in the pump-and-probe measurements. In a
photoreflectance pump-probe measurement, the change of the reflectance or
the change of polarization of probe-laser pulses reflected from the sample
surfaces is detected with variable time delay against the excitation of the
sample by the pump pulse [30]. Since the coherent phonons induce the change
of the dielectric constant or linear susceptibility, coherent phonons are ob-
served as the oscillatory behavior of the intensity or the polarization of the
reflected probe laser light, whose oscillation frequency and the damping rate
are directly related to the frequency and relaxation rate of coherent lattice
vibrations, respectively. In such pump-probe measurements, all the observed
phonon modes were Raman-active, and no Raman-inactive4 phonon modes
have been observed so far. This fact suggests that the Raman activity is es-
sential for the excitation and detection of the coherent phonons observed in
the optical pump-probe measurements. Therefore, the impulsive-stimulated
Raman scattering is widely accepted as the excitation mechanism for the
coherent phonons [31, 32, 33]. However, the Raman-scattering process does
4
Note that an optical phonon mode can be both infrared-active and Raman-active
when the crystal does not have the inversion symmetry, depending on the mode
symmetry.
80 Ping Gu and Masahiko Tani

not emit electromagnetic radiation at THz frequencies. Therefore, the exci-


tation mechanism for coherent phonons that emit THz radiation needs to be
sought in other processes, such as the ultrafast surface-field screening and
the ultrafast build-up of the photo-Dember field, as discussed above.
In the first part of this section we describe THz radiation from coherent
phonons in semiconductors with several experimental results obtained by the
authors. In the latter part of this section, we describe THz radiation from
coherent plasma oscillations and coupled oscillations of coherent phonon and
plasmon excited in semiconductors.

3.2 THz Radiation from Coherent Phonons


3.2.1 THz Radiation from Coherent Phonons in Telluride
Compound Semiconductors
The first observation of the THz emission from coherent phonons was re-
ported by Dekorsy et al. for an LO phonon of an only-infrared(IR)-active
mode (A2 -mode) in Te [10, 22]. Similar results were also reported by Tani
et al. [34] for PbTe and CdTe, whose optical phonon modes are only-IR-active
and IR-and-Raman-active, respectively. In the following, we review the exper-
imental results obtained by Tani et al. [34] for the three semiconductors, that
is, Te, PbTe and CdTe. These telluride semiconductors have low-frequency
optical phonons due to the heavy mass of tellurium atom and are appropriate
for observation of THz emission from coherent phonons using detectors with
a limited detection bandwidth, such as a PC antenna (< 6 THz).

3.2.2 Experimental Setup


They used a mode-locked Ti: sapphire laser, whose wavelength, pulse width,
and repetition rate were ≈ 800 nm, ≈ 70 fs and 82 MHz, respectively. Laser
pulses were irradiated onto the sample surface at an incident angle of 45◦
and was loosely focused to a diameter about 1 mm. The photoexcited car-
rier densities for Te, PbTe, and CdTe were estimated as 5.2 ×1017 cm−3 ,
3.3 ×1017 cm−3 , and 1.2 ×1016 cm−3 , respectively, by using the absorption
depth of the pump laser (12 nm, 18 nm, and 0.5 µm for nondoped Te, PbTe,
and CdTe, respectively). Because these photocarrier densities are not so high
that the phonon–plasmon coupling effect becomes significant, the THz emis-
sion they observed for undoped samples should thus be related to almost
pure LO-phonon modes. The radiation emitted in the direction of the optical
reflection was collected and focused onto a PC sampling detector by a pair
of parabolic mirrors.

3.2.3 Sample Properties


The properties of samples used by Tani et al. are listed in Table 2. All
samples were monocrystalline. The sample of Te had a surface perpendic-
Terahertz Radiation from Semiconductor Surfaces 81

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]

Table 2. Sample properties

Sample Crystal Optical-phonon mode Raman IR


structure activity activity
Te hexagonal A1 = 3.6 THz yes no
(c-axis ⊥ surface) A2 : TO/LO = 2.6/2.82 THz no yes
E  : TO/LO = 2.76/3.09 THz yes yes
E  : TO/LO = 4.22/4.26 THz yes yes
PbTe (100) rock-salt TO/LO = 0.96/3.42 THz no yes
CdTe (100) zincblende TO/LO = 4.20/5.08 THz yes yes

3.2.4 Spectra of Te, PbTe, and CdTe

Figures 11a–c show the time-domain signal waveforms detected by the PC


sampling detector for THz radiation from Te, PbTe, and CdTe surfaces, re-
spectively. For all the samples the radiation was p-polarized, suggesting that
it was emitted by the electric polarization perpendicular to the sample sur-
face. Figures 12a–c show the Fourier-transformed amplitude spectrum of each
waveform in Fig. 11.
Te In Fig. 11a, the first cycle of the oscillations corresponds to the ra-
diation due to the transient electric polarization caused by the ul-
trafast build-up of the photo-Dember field at Te surface. The subse-
quent oscillations persisting for about 3 ps correspond to THz radi-
ation from the coherent LO-phonon of the A2 -mode in Te [10, 22].
The Fourier-transformed amplitude spectrum shows a spectral peak
at 2.83 THz ± 0.05 THz with a FWHM width of 0.4 THz (Fig. 12a).
No coherent radiation with a frequency corresponding to the TO-
phonon mode is observed. Instead, there is a spectral dip at the TO-
phonon frequency (2.6 THz). Radiation from the other IR-active modes
(E  fTO/LO and ETO/LO

), whose polarization axes are parallel to the
82 Ping Gu and Masahiko Tani

Fig. 11. Differential sig-


nals detected by the PC di-
pole antenna for the emission
from (a) Te, (b) PbTe, and
(c) CdTe surfaces

surface (perpendicular to the c-axis), were not observed since these


modes do not couple to the field normal to the surface induced by the
photo-Dember effect.
PbTe For PbTe, strongly damped oscillations (Fig. 11b) have been observed
without significant initial radiation burst, which is observed for the
other two samples. The Fourier-transformed spectrum (Fig. 12b) shows
a broad spectral peak located at 3.2 THz ± 0.1 THz and having a
1.3 THz bandwidth (FWHM). This coherent radiation is attributed
to the emission from the LO-phonons excited in PbTe (3.4 THz). The
slight shift to the lower frequency may be due to a rapid decrease of
the detector responsivity at higher frequencies. As is the case of the
only-IR-active A2 mode in Te, any coherent phonon oscillation was
Terahertz Radiation from Semiconductor Surfaces 83

Fig. 12. Fourier-transformed


amplitude spectra for the sig-
nal waveforms in Figs. 11a–c.
The TO- and LO-phonon fre-
quencies of each sample are
indicated by vertical arrows.
The calculated emission ef-
1/2
ficiencies Ga for radiation
from PbTe and CdTe into the
air are indicated by dashed
lines. The dashed-dotted line
in Fig. 12c is the spectrum for
CdTe magnified by a factor of
10 in order to show the struc-
ture at higher frequencies

not observed in the transient reflectivity measurement for the undoped


bulk PbTe5 .
The driving force for the excitation of the coherent phonons in PbTe,
as in the case of Te, can be attributed to the ultrafast build-up of
the photo-Dember field [22]. Since the optical phonons in PbTe are
Raman-inactive, the Raman process is excluded from the excitation
mechanism for the coherent phonons. The displacive excitation mecha-
nism [36] is also excluded since this excitation scheme is applicable only
to totally symmetric modes such as A1 mode. Since the surface built-in
5
Oscillations in the transient reflectivity signal corresponding to the LO-phonon
frequency were observed for a heavily doped n-PbTe (n > 1019 cm−3 ), for which
the parity selection rule is expected to be broken by the high surface field or
stress introduced by the high-density carriers, giving rise to the Raman activity.
84 Ping Gu and Masahiko Tani

field for PbTe is expected to be weak because of the small bandgap


energy (Eg = 0.29 eV) and the large dielectric constant (εs = 388),
the instantaneous screening of the surface field by the photoexcited
carriers [37] cannot be a strong driving force for excitation of coherent
phonons in PbTe. On the other hand, from the large inhomogeneity in
photocarrier density due to the small absorption length and the large
diffusivity of the highly energetic photoexcited carriers (with excess en-
ergy of ∆E = 1.26 eV), a strong and fast build-up of the photo-Dember
field normal to the surface is expected, as in the case of Te [22].
CdTe For CdTe, fast oscillations after the strong radiation burst are observ-
ed (Fig. 11c). The Fourier-transformed amplitude spectrum (Fig. 12c)
shows a spectral peak at 5.1 THz ± 0.05 THz with a FWHM width
of 0.51 THz. This radiation component is attributed to the emission
by coherent LO-phonons in CdTe. There is a spectral dip at the TO-
phonon frequency (4.2 THz) as well. Because the detector responsivity
is low around 5 THz the amplitude of the radiation from the coherent
LO-phonon is estimated to be as large as the amplitude of the first
radiation burst. As expected from the Raman activity of the optical
phonon modes in CdTe, oscillations of both the coherent TO- and LO-
phonons were observed in the transient photoreflectance measurement
(not shown). As in other polar semiconductors with wide bandgaps
(> 0.4 eV), such as GaAs or InP, the excitation of coherent phonons
in CdTe can be attributed to the ultrafast screening of the surface
built-in electric field by the photocarriers [37].

3.2.5 Spectral Dip at TO-Phonon Frequency

In addition to the pronounced coherent emission at the LO-phonon frequency,


a striking feature persistent in the observed spectra of THz radiation shown
above is the almost complete lack of spectral components at the TO-phonon
frequencies. These spectral dips at the TO-photon frequencies are not well
explained by the absorption at TO-phonon frequencies alone.
Although coherent emission from pure TO-phonon oscillations is not prob-
able, coherent THz emission near the TO-phonon frequency is expected from
the lower-frequency mode of the LO-phonon–plasmon (L− mode) at carrier
densities higher than 1018 cm−3 [38]. (The details of the LO-phonon–plasmon
coupled modes are discussed in Sect. 3.3). Such an LO-phonon–plasmon
coupled mode has been actually observed for n-GaAs in the optical tran-
sient reflectivity measurements [39, 40], which showed LO-phonon–plasmon
coupling-mode frequencies determined by the total carrier density including
the contribution from both the photoexcited carriers and the residual carri-
ers by doping. The total lack of the emission spectrum near the TO-phonon
frequency or L− mode thus contradicts the observation of these modes in the
optical transient reflectivity measurements. The reason may be attributed to
the emission property of the coherent phonons from the sample surfaces.
Terahertz Radiation from Semiconductor Surfaces 85

Fig. 13. Geometry for calculations of the radi-


ation pattern in air from a dipole on a dielec-
tric substrate using the reciprocity theorem.
Ei is the incident electric field from element
I2 , θa is the incident angle and θd is the re-
fracted angle

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.

3.2.6 Radiation Pattern and Emission Efficiency

Let us consider the radiation pattern of THz radiation from a semiconductor


surface excited by a femtosecond laser pulse in order to clarify the mechanism
of the spectral dip observed at the TO-phonon frequency.
It is well known that an emitting antenna on a thick dielectric substrate
emits most of its radiation to the dielectric side, rather than to the air side.
The ratio of the total emitted power into the dielectric, Pd , to that into
air, Pa , is approximated by (1/2)εsd, where εd is the dielectric constant of
the substrate and s is a constant depending on the type of antenna and its
orientation. For example, s equals 3/2 for an infinitesimal dipole antenna
parallel to the dielectric surface [41]. Since, in the frequency region near
optical phonons, the value of the dielectric constant changes significantly
with frequency, the fraction of emitted radiation to free space, Pa /(Pa + Pd ),
also changes significantly with frequency.
In the present case, the radiation source is the transient electric polariza-
tion at the semiconductor surface and is considered to be distributed infin-
itesimal dipoles (Hertzian dipoles) with their polarization perpendicular to
the surface and the relative phase among them determined by the incidence
of the optical wavefront of the pump-laser pulses. In this case, most of the
radiation is emitted in the direction determined by the generalized Snell’s
law given by (9), which we rewrite for an incident angle of the pump laser,
86 Ping Gu and Masahiko Tani

θi ,

sin θi = sin θa =  [ñd (ω)] sin θd = nd (ω) sin θd , (14)

where θa (= θi ) is the direction (polar angle) of the radiation emitted into


free space, θd is the direction of the radiation emitted into the dielectric,
ñd (ω) = nd (ω) − iκd (ω) = εd (ω) is the complex refractive index of the
dielectric (semiconductor) at the radiation frequency ω (Fig. 13).
All other emissions except the emission in directions to θa and θd become
null because of the destructive interference between the field components
from the distributed dipoles. The effect of diffraction is not expected to be
significant when the excitation spot at the surface is large enough compared
to the wavelength of the relevant THz radiation. For the present case, the
pump-laser spot on the semiconductor surface was about 1 mm, which is large
enough for the wavelength of the relevant radiation (< 0.3 mm).
The far-field radiation pattern of an elementary antenna on a dielectric
surface can be calculated easily by using the Lorentz reciprocity theorem [41],
by which we can avoid having to solve the complex boundary-value problem
directly. To obtain the radiation pattern of a dipole (dipole 1) on a dielectric
surface that is oriented perpendicular to the surface, we assume a test dipole
(dipole 2) in air, as depicted in Fig. 13, in which a current element I1 in the
dipole 1 produces a field component E1 parallel to a current element I2 in the
test dipole 2 and the current element I2 in turn produces a field component
E2 parallel to I1 . According to the Lorentz reciprocity theorem, if I1 = I2 ,
then E1 = E2 . Because we know the radiation field of dipole 2, we can easily
calculate E2 , which is the transmitted field component parallel to I1 of the
incident field on the interface, Ei . We can then write

E1 = E2 = t(p) sin θd Ei , (15)

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

radiation pattern to the air-side half-space for an infinitesimal dipole placed


just beneath the dielectric surface and oriented to the surface normal. Similar
consideration for the dielectric side gives
3 n3d |ñd |2 sin2 θd cos2 θd 3
Pd (θd ) = P0 ≡ P0 gd (θd ) . (18)
8π |ñd cos θa + cos θd |2 8π
Here, gd (θd ) represents the radiation pattern to the dielectric-side half-space
for an infinitesimal dipole placed just beneath the dielectric surface and ori-
ented to the surface normal.
The above equations for Pa (θa ) and Pd (θd ) are valid for an element dipole
at the air/dielectric interface. The radiation pattern of the distributed dipoles
excited with a pump laser at an incident angle, θi , is given by using the
principle of pattern multiplication in the conventional antenna theory, which
states that the electric field or intensity pattern of an array consisting of
similar elements is the product of the pattern of a single-element antenna
(the element pattern) and the pattern of an array of isotropic point sources
with the same locations, relative amplitudes and phases as the original array
(the array factor) [42].
In the present case, the element pattern is the radiation pattern, ga (θ) and
gd (θ), given by (17) and (18). The array factor, f (θ, φ), in this case is equal
to the optical beam pattern reflected or refracted at the semiconductor/air
interface when we neglect the diffraction effect for THz radiation, and is
approximated by a delta function centered at spherical angles (θ = θa , φ = 0)
and (θ = θd , φ = 0), where the polar angle θ is defined as the angle from the
surface normal and the azimuthal angle φ is defined as the angle from the
reflection plane. In the present experimental geometry, θa = θi = 45◦ and
θd (ω) = arcsin [sin 45◦ /nd (ω)].
Thus, the ratio of the radiation power emitted into free space to the total
emitted radiation power is approximately given by

air fa (θ, φ)ga (θ) dΩ


Ga (ω) =
f
air a
(θ, φ)g a (θ) dΩ + diel fd (θ, φ)gd (θ) dΩ

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

where ε∞ and εs denote the high-frequency and static dielectric constants,


ωTO is the TO-phonon frequency, and Γ is the phonon damping parameter.
Although the contribution from free carriers (the Drude term) is neglected
in (21), the inclusion of carrier effects only smears out the TO-phonon reso-
nant dispersion of εd (ω) and hence Ga (ω) as well.
We carried out the calculations of Ga (ω) for PbTe, with ε∞ = 32.8,
εs = 388, and Γ = 26 cm−1 [43] and for CdTe with ε∞ = 7.1, εs = 10.2,
and Γ = 8.7 cm−1 [44]. The square root of the calculated Ga (ω) is shown by
the dashed lines in Fig. 12. The curves show a peak around the LO-phonon
frequency, where the dielectric constant, |εd | = n2d + κ2d , takes a minimum,
1/2
and a dip or a reduction of Ga (ω) around the TO-phonon frequency, where
the dielectric constant takes a maximum. Thus, the enhanced amplitude of
radiation from the coherent LO-phonons is explained by the increased out-
coupling efficiency of the radiation from the crystal to free space due to
the small dielectric constant around the LO-phonon frequency. The spectral
dip at the TO-phonon frequency is explained by the strong reduction of
the outcoupling efficiency [Ga (ωTO ) 1] due to the large dielectric constant
around the TO-phonon frequency. The lack of an initial radiation burst from
PbTe is also explained by a significantly low outcoupling efficiency over a
range of frequency below the TO-phonon frequency (< 1 THz), resulting in
suppression of the emission of radiation by the photo-Dember field. Although
the calculation was not carried out for Te because of the complication due
to the presence of many optical phonon modes, the same considerations also
apply to the spectral dip at the TO-phonon frequency of the A2 -mode of Te,
where the dielectric constant takes a large value (|εd | > 100) [45].

3.3 Coupling of LO Phonon and Plasmon

Now we discuss THz radiation from plasma oscillations coupled to coherent


phonon oscillations. Excitation by the femtosecond laser pulses drives not
only the coherent lattice vibrations but also the collective motion of carriers,
that is, the plasmon oscillations in semiconductors. In the absence of coupling,
Terahertz Radiation from Semiconductor Surfaces 89

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

the plasmon is expected to oscillate with the plasma frequency as mentioned


earlier:

ωp = 4πe2 N/εε0 m∗ , (22)

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

effect of the carriers. Note that although the frequency of L− mode, ω− , is


near the TO-phonon frequency, it is not a transverse mode but a longitudinal
mode of lattice vibrations.
The frequencies of the LO-phonon–plasmon coupled modes are gener-
ally carrier-density dependent. However, in the transient reflectance mea-
surements we usually observe density-independent oscillations at TO-phonon
and LO-phonon frequencies: at low carrier density the oscillations at the LO-
phonon frequency are dominant, while at high carrier density the oscillations
at the TO-phonon frequency are dominant.
Kuznetsov and Stanton [38] investigated the LO-phonon–plasmon inter-
action in detail and made a theoretical calculation on GaAs. They included
the effect of inhomogeneous density distribution in the calculation and found
that the oscillations near the TO-phonon and the LO-phonon frequencies
are preserved in the averaging process over different carrier densities due to
Terahertz Radiation from Semiconductor Surfaces 91

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.

3.3.1 InSb Phonon–Plasmon THz Emission

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. 16. Time-domain wave-


forms of the THz radiation
detected in (a) reflection and
(b) transmission geometry for
the same InSb film. The inset
is the waveform of (a) magni-
fied by a factor of 3 at a delay
time of more than 3.5 ps

in Fig. 17 with a magnified scale. By fitting a Lorentzian function to these


spectral peak profiles, the second spectral peak in Fig. 17a and b is found at
3.94 THz ± 0.05 THz, and the third peak in Fig. 17a at 6.10 THz ± 0.05 THz,
respectively.
No significant coherent radiation was observed at the bare TO- and LO-
phonon frequencies (indicated by vertical arrows in Fig. 17a). The measured
frequencies of 3.94 THz and 6.10 THz agreed well with those of the lower
(L− ) and upper (L+ ) branches of the phonon–plasmon coupled modes ex-
pected from the theoretical calculation with the doped carrier density of
(6.70 ×1016 )cm−3 . These results indicate that the coherent radiation we ob-
served originates from L− and L+ modes. The absence of L+ mode radiation
in the transmission geometry is due to the strong absorption in the 0.7 mm
thick GaAs substrates: the internal absorption is estimated to be over 90%
at 6.10 THz and less than 40% at 3.94 THz in the GaAs substrate.
In Fig. 17a, the magnitude of the THz-radiation spectrum at the L− mode
frequency (3.94 THz) is almost half that at 1 THz where the surge-current
contribution is dominant. The sensitivity of a PC-antenna detector decreases
rapidly with increase of frequency: The sensitivity of the PC-antenna detector
at 4 THz and 6 THz is, respectively, estimated to be one-fifth and one-fifteenth
Terahertz Radiation from Semiconductor Surfaces 93

Fig. 17. (a) and (b) are the Fourier-


transformed spectra of the data in
Figs. 16a and b. The broken lines in
(a) and (b) are the frequency spectra
of L− and L+ modes magnified by a
factor of 2. The frequencies of the TO-
and LO-phonons, and the L+ and L−
modes are denoted by vertical arrows

of that at 1 THz. Therefore, the spectral intensities due to the LO-phonon–


plasmon coupling modes are actually very intense.
Figure 18 shows the THz-radiation spectra obtained with pump laser
powers of 20 mW (photocarrier density nex ≈ (2.5 × 1016 ) cm−3 ), 40 mW
(nex ≈ (5.0 × 1016 ) cm−3 ), and 60 mW (nex ≈ (7.5 × 1016 ) cm−3 ) in the
(a) reflection geometry and (b) transmission geometry, for a limited win-
dow to remove the contribution of the THz radiation component due to the
surface surge current. Figures 18a and b show that the peak frequencies of
the coherent THz radiation corresponding to the L− and L+ modes do not
change with the excitation power, while the spectral amplitudes of the these
modes increase with the pump power. This is rather surprising when we sup-
pose that the L− and L+ modes frequencies are determined by the total
carrier density (the sum of the intrinsic and photoexcited carriers). This re-
sult clearly indicates that the oscillation frequencies of the phonon–plasmon
coupling modes depend only on the intrinsic carrier concentration and not
on the density of photogenerated carriers. Similar results were reported by
Kersting et al. [47] for the coherent THz radiation from plasma oscillations
in GaAs, the frequency of which was independent of the photogenerated car-
rier density. In addition, the decay time of oscillations for the L− mode does
not change with excitation power. This means that the dephasing of the
L− modes is not affected by the optically excited carrier dynamics such as
94 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

carrier–carrier scattering. A similar result was obtained for the L+ mode,


too.
On the other hand, a time-domain optical pump-probe measurement [48]
and CW-Raman scattering measurement [49] have shown that the photogen-
erated carriers affect the frequency and damping time of the coupling modes.
The origin of this discrepancy is not clear at present, and further investiga-
tions are needed.

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

field amplitude is proportional to the carrier acceleration, information on


the ultrafast carrier transport such as the peak velocity in the overshoot
regime and the influence of the lattice vibrations are obtained from the time-
resolved measurement of the emitted radiation (with a proper correction for
the detector response). For example, Leitenstorfer et al. [50] measured the
time-resolved THz field amplitude from InP and GaAs surface with a time
resolution of 20 fs. They measured electron peak velocities of (6 × 107 ) cm/s
and (8 × 107 ) cm/s for GaAs and InP, respectively, under an electric field of
90 kV/cm. These overshoot velocities are much higher than the drift veloci-
ties in the steady-state regime, and the one for InP was larger than that of
GaAs, reflecting the larger energy separation between the central valley and
the side minima in the InP conduction band compared to that in GaAs.
Optical phonon dynamics are also investigated by measuring the THz ra-
diation from the polar semiconductor surface because the carrier transport
is influenced by the lattice vibration or the electron plasma coupled with
LO-phonons. THz radiation from coherent phonons or coherent plasmons are
interesting phenomena, which are still not well understood. For example, the
emission frequencies for L− and L+ modes are dependent only on the intrin-
sic carrier density, and are not dependent on the excitation carrier density.
The radiation power from coherent phonons and plasmon are quite efficient,
considering the low sensitivity at high frequencies of the PC antennas used as
the detector, which contradict the theoretical prediction by Kuznetsov and
Stanton [38].

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

[12] B. Greene, P. Saeta, D. Dykaar, S. Schmitt-Rink, S. Chuang: IEEE J. Quantum


Electron. 28, 2302 (1992) 63, 64, 66
[13] S. Howells, L. Schlie: Appl. Phys. Lett. 67, 3688 (1995) 63
[14] X.-C. Zhang, Y. Jin, T. Hewitt, T. Sangsiri, L. Kingsley, M. Welner: Appl.
Phys. Lett. 63, 2003 (1993) 63
[15] N. Sarukura, H.Ohtake, S. Izamida, Z. Liu: J. Appl. Phys. 84, 1 (1998) ac-
cording to a recalibration of the bolometer sensitivity by Sarukura et al., the
total radiation power from an InAs surface under a magnetic field of 1 T is
corrected to be about 50 µW with a pump power of 1 W. 1398 (2000) I.; P.N.
Saeta, D.R. Dykaar, S. Schmitt-Rink, and S.L. Chuang, IEEE J. Quantum
Electron. 28, 2302 (1992) 63, 71
[16] A. Bonvalet, M. Joffre, J. Martin, A. Migus: Appl. Phys. Lett. 67, 2907 (1995)
66
[17] A. Rice, Y. Jin, X. Ma, X.-C. Zhang, D. Bliss, J. Larkin, M. Alexander: Appl.
Phys. Lett. 64, 1324 (1994) 66
[18] X.-C. Zhang, Y. Jin, K. W. X. Ma, A. Rice, D. Bliss, J. Larkin, M. Alexander:
Appl. Phys. Lett. 64, 622 (1994) 66
[19] P. Gu, M. Tani, S. Kono, K. Sakai: J. Appl. Phys. 91, 5533 (2002) 68, 91
[20] S. Singh: Handbook of Lasers (CRC, Cleveland 1971) 69
[21] For review papers, see, for example, X.-C. Zhang and D.H. Auston: J. Appl.
Phys. 71, 326 (1992), S.C. Howells and L.A. Schlie: Appl. Phys. Lett. 67, 3688
(1995), T. Kondo, M. Sakamoto, M. Tonouchi, and M. Hangyo: Jpn. J. Appl.
Phys. 38, L1035 (1999), M. Hangyo, M. Migita, and K. Nakayama: J. Appl.
Phys. 90, 3409 (2001), M.B. Johnston, D.M. Whittaker, A. Corchia, A.G.
Davies, and E.H. Linfield: J. Appl. Phys. 91, 2104 ( 2002), M.B. Johnston,
D.M. Whittaker, A. Corchia, A.G. Davies, and E.H. Linfield: Phys. Rev. B
65, 165301 (2002) 70
[22] T. Dekorsy, H. Auer, H. Bakker, H. Roskos, H. Kurz: Phys. Rev. B 53, 4005
(1996) 71, 78, 80, 81, 83, 84
[23] S. Kono, P. Gu, M. Tani, K. Sakai: Appl. Phys. B 71, 901 (2000) 72
[24] W. Mönch: Semiconductor Surface and Interface (Springer, Berlin, Heidelberg
1993) p. 68 73
[25] S. Howells, S. Herrera, L. Schlie: Appl. Phys. Lett. 65, 2946 (1994) 76
[26] H. Takahashi, Y. Suzuki, M. Sakai, A. Ono, N. Sarukura, T. Sugiura, T. Hi-
rosumi, M. Yoshida: Appl. Phys. Lett. 82, 2005 (2003) 76
[27] M. Nuss, D. Auston, F. Capasso: Phys. Rev. Lett. 58, 2355 (1987) 76
[28] P. Y. Yu, M. Cardona: Fundamentals of Semiconductors, 3rd ed. (Springer,
Berlin, Heidelberg 2001) pp. 110–113 77
[29] Coherent THz emission in semiconductors, in K. Tsen (Ed.): Ultrafast Physical
Properties in Semiconductors, Semicond. Semimet. 67 (Academic, New York
2001) 79
[30] For the pump-probe experiment of coherent phonons, see for example,
W. Kütt: Adv. Solid State Phys. 32, 113 (1992), W. Kütt, W. Albrecht, and
H. Kurz: IEEE J. Quantum Electron. 28, 2434 (1992), H.J. Zeiger, J. Vidal,
T.K. Cheng, E.P. Ippen, G. Dresselhaus, and M.S. Dresselhaus: Phys. Rev. B
45, 768 (1992) 79
[31] Y.-X. Yan, J. E.B. Gamble, K. Nelson: J. Chem. Phys. 83, 5391 (1985) 79
[32] S. Ruhman, A. Joly, K. Nelson: J. Chem. Phys. 86, 6563 (1987) 79
Terahertz Radiation from Semiconductor Surfaces 97

[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

cadmium telluride (CdTe), 81, 84 indium phosphide (InP), 65–66, 69–73


coherent LO phonon, 77–78
lead telluride (PbTe), 80–85, 88
difference frequency mixing (DFM), 64, LO phonon–plasmon mode, 79, 91
67
optical rectification, 64, 66, 68–69
gallium arsenide (GaAs), 65, 70, 89, 90
photo-Dember effect, 64, 70–74, 76
impulsive-stimulated Raman scattering, plasmon, 78–80, 88–90
79 plasmon frequency, 89
indium antimonide (InSb), 64–67,
69–71, 73–76, 91–92 surge current, 64, 70
indium arsenide (InAs), 63–67, 69–71,
73–76 tellurium (Te), 65, 80–82

You might also like