[Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and...

42
Terahertz Radiation from Bulk and Quantum Semiconductor Structures Yutaka Kadoya 1 and Kazuhiko Hirakawa 2 1 Department of Quantum Matter, ADSM, Hiroshima University Higashi-hiroshima 739-8530, Japan [email protected] 2 Institute of Industrial Science, The University of Tokyo Komaba, Meguro-ku 153-8505, Japan [email protected] Abstract. An excitation of various semiconductor structures by femtosecond opti- cal pulses leads to radiation of electromagnetic waves in the THz-frequency range. Since the waveform reflects the dynamics of the excited carriers or polarizations from which the THz wave is radiated, the time-domain studies of the emitted THz wave provide a unique way of looking directly into the temporal as well as spatial evolution of the excitation on the subpicosecond time scale. In this Chapter, we present recent studies on the emission of THz waves from nonstationary carrier transport and the coherent oscillation of carrier population in bulk and quantum semiconductor structures. Also discussed in this Chapter are the possibilities of the use of semiconductor structures for the improvement or the development of THz-wave emitters. 1 Introduction As described in other chapters, impulsive currents in semiconductors excited by femtosecond (fs) optical pulses radiate electromagnetic (EM) waves in the THz-frequency range, serving, for example, as an EM wave source for far- infrared spectroscopy. The EM wave in the far field is proportional to the time derivative of the current, ∂J/∂t. In a simplified view, J can be expressed as J = eNv, where e, N , and v are the elementary electric charge, total number of photoexcited carriers (electron or hole) and their velocity, respectively. If the carriers are excited by sufficiently short laser pulses and their lifetime is long enough, N can be regarded as being constant and we may approximate as ∂J/∂t ∂v/∂t. Hence, in principle, the observation of such THz EM waves provides us with a way of looking into the temporal variation of v, namely the acceleration/deceleration dynamics of photoexcited carriers [1]. Actually, there is no other useful technique that can detect the carrier transport on a subpicosecond time scale. In the case of off-the-shelf semiconductor wafers, however, the acceleration is caused by the DC electric field, which is built- in near the surfaces and is out of control. Consequently, unwanted effects such as the photocarrier-induced screening make quantitative discussion quite difficult [2, 3]. K. Sakai (Ed.): Terahertz Optoelectronics, Topics Appl. Phys. 97, 117–158 (2005) © Springer-Verlag Berlin Heidelberg 2005

Transcript of [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and...

Page 1: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and QuantumSemiconductor Structures

Yutaka Kadoya1 and Kazuhiko Hirakawa2

1 Department of Quantum Matter, ADSM, Hiroshima UniversityHigashi-hiroshima 739-8530, [email protected]

2 Institute of Industrial Science, The University of TokyoKomaba, Meguro-ku 153-8505, [email protected]

Abstract. An excitation of various semiconductor structures by femtosecond opti-cal pulses leads to radiation of electromagnetic waves in the THz-frequency range.Since the waveform reflects the dynamics of the excited carriers or polarizationsfrom which the THz wave is radiated, the time-domain studies of the emitted THzwave provide a unique way of looking directly into the temporal as well as spatialevolution of the excitation on the subpicosecond time scale. In this Chapter, wepresent recent studies on the emission of THz waves from nonstationary carriertransport and the coherent oscillation of carrier population in bulk and quantumsemiconductor structures. Also discussed in this Chapter are the possibilities ofthe use of semiconductor structures for the improvement or the development ofTHz-wave emitters.

1 Introduction

As described in other chapters, impulsive currents in semiconductors excitedby femtosecond (fs) optical pulses radiate electromagnetic (EM) waves inthe THz-frequency range, serving, for example, as an EM wave source for far-infrared spectroscopy. The EM wave in the far field is proportional to the timederivative of the current, ∂J/∂t. In a simplified view, J can be expressed asJ = eNv, where e, N , and v are the elementary electric charge, total numberof photoexcited carriers (electron or hole) and their velocity, respectively. Ifthe carriers are excited by sufficiently short laser pulses and their lifetime islong enough, N can be regarded as being constant and we may approximateas ∂J/∂t ∝ ∂v/∂t. Hence, in principle, the observation of such THz EM wavesprovides us with a way of looking into the temporal variation of v, namelythe acceleration/deceleration dynamics of photoexcited carriers [1]. Actually,there is no other useful technique that can detect the carrier transport on asubpicosecond time scale. In the case of off-the-shelf semiconductor wafers,however, the acceleration is caused by the DC electric field, which is built-in near the surfaces and is out of control. Consequently, unwanted effectssuch as the photocarrier-induced screening make quantitative discussion quitedifficult [2, 3].

K. Sakai (Ed.): Terahertz Optoelectronics, Topics Appl. Phys. 97, 117–158 (2005)© Springer-Verlag Berlin Heidelberg 2005

Page 2: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

118 Yutaka Kadoya and Kazuhiko Hirakawa

By virtue of the well-developed thin-film crystal growth techniques ofthe present day, various types of layered semiconductor structures can befabricated quite precisely and reproducibly. A simple one is the undoped(∼ intrinsic) layer grown on a doped substrate (i-n or i-p structures). Insuch structures, contrary to off-the-shelf semiconductors, the DC field canbe easily controlled with the external bias. Therefore, the use of artificiallystructured samples is important even in the study of carrier dynamics inbulk semiconductors for a quantitative discussion. Section 2, in this Chapter,presents a rigorous elucidation of nonstationary carrier transport in bulksemiconductors.

Another important class of artificial semiconductor structures is, needlessto say, the quantum structures, such as quantum wells (QWs) and superlat-tices (SLs). The excitation of these structures by fs optical pulses also triggersthe radiation of THz EM waves, providing rich information on quantum-mechanical phenomena such as the tunneling of the electron wave packet andthe coherent evolution of the excitation. The radiation of THz waves fromsuch structures can be viewed classically in two ways. One is the current, J ,flowing across the heterostructures, as in the case of bulk semiconductors.The other is the transient or oscillating polarization, P , induced by the shortoptical pulses. Although, under the local electric dipole approximation bothviews are related by J = ∂P/∂t, the description based on the polarizationis suited when the coherence of the excitation plays an important role. Inthe case of the wave-packet motion, for example, the polarization arises fromthe different time evolution of the wavefunctions between the electrons andholes.

Historically, the radiation of THz waves from the wave-packet motionswas observed first in coupled [4] and single [5] QWs in 1992, and then inSLs (Bloch oscillation) in 1993 [6]. Since then, a good number of reportshave been published. In contrast to widely used nonlinear spectroscopic tech-niques such as four-wave mixing, where in general the interpretation of thesignal is not simple, the THz signals reflect rather directly the dynamics ofthe carrier motion or population, giving a big advantage to the THz-basedtechnique. Indeed, for example, the theoretical frameworks, which are usedfor the analysis of the nonlinear optical response, have been critically testedon the basis of the THz-emission experiments [7]. In this Chapter, two recentTHz-emission studies on semiconductor quantum structures will be presentedin detail; one on the Bloch oscillation and the miniband transport in SLs inSect. 3 and the other on the coherent nutation of the exciton populationinherent to the strongly coupled exciton–photon system in Sect. 4.

On the other hand, the use of semiconductor structures is expected toopen up a possibility of improving or developing THz emitters, detectorsand others, which may be desired in future high-performance THz-imagingand/or spectroscopy systems. As examples, the amplification or generationof THz waves based on the negative differential resistance in SLs and the en-

Page 3: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 119

hancement of radiation from semiconductor surfaces by the use of microcavitystructures will be discussed in Sects. 3 and 5, respectively.

2 Nonstationary Carrier Transport in Strongly BiasedBulk Semiconductors

2.1 A Brief Review on Nonstationary Carrier Transportand THz Wave Radiation

With the recent progress in miniaturization of high-speed transistors, ultra-fast transistors with cutoff frequencies well above 500 GHz have been real-ized [8]. In such ultrashort channel transistors, carriers experience very fewscattering events in the channel and drift in a very nonstationary manner,as schematically illustrated in Fig. 1. Consequently, the performance of suchultrafast transistors is not mainly determined by the steady-state properties,such as saturation velocities and mobilities, but is governed by the nonsta-tionary carrier transport subjected to high electric fields. However, it has beendifficult to characterize such very fast phenomena by using conventional elec-tronics, such as sampling oscilloscopes, because of their limited bandwidth.Consequently, Monte Carlo calculations have been the only tool for discussingtransient carrier transport.

Very recently, the pioneering work by Leitenstorfer et al. [9,10] has openedup a way to characterize the transient carrier velocities with sub-100 fs timeresolution by measuring the THz EM radiation emitted by photoexcited car-riers in semiconductors. According to Maxwell’s equations, when chargedcarriers are accelerated by electric fields, they emit radiation whose electric-field component, ETHz, is proportional to the acceleration, ∂v/∂t. Since theacceleration/deceleration of carriers in semiconductor devices takes place onthe subpicosecond time scale, the emitted radiation is in the THz range. Thebasic idea is that transient carrier velocities can be recovered by integrat-ing the traces of detected ETHz with respect to time. By using this scheme,Leitenstorfer et al. [9,10] successfully determined the transient carrier veloci-ties in GaAs and InP, and opened up the possibility of quantitative discussionof velocity-overshoot behaviors.

What is represented by the detected THz signals is, however, not fullyunderstood yet. Although Leitenstorfer et al. already pointed out that thefinite sample thickness distorts the apparent transient-velocity characteris-tics, quantitative discussion of such an effect has not been carried out yet.Another important point is that the velocities of electrons and holes cannotbe separated in the photoexcited THz measurements; in general, it is be-lieved that the THz signal observed experimentally is dominated by electrontransport because of its lighter effective mass. However, it is necessary toexamine whether the hole contribution is indeed negligible or not. Therefore,it is crucial to make a quantitative comparison between the nonequilibrium

Page 4: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

120 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 1. Carrier transport in long diffusive channels (a) and very short ballisticchannels (b). In conventional long-channel transistors (a), their performance isdetermined by steady-state properties, such as saturation velocities and mobilities.However, in very short channel transistors (b), carrier transport is quasiballisticand their switching speed is governed by how fast carriers are accelerated

Fig. 2. (a) Sample structure and (b) schematic illustration of THz-emission ex-periment on biased bulk GaAs

transport data obtained by the time-domain THz spectroscopy and ensembleMonte Carlo (EMC) results to see how quantitative discussions can be madefrom THz data [11].

2.2 Time-Domain THz Measurements of Transient CarrierVelocities in Strongly Biased Semiconductors

The sample used in this section consisted of a 1µm thick undoped GaAsgrown on n+-GaAs substrate (see Fig. 2a). In this experiment, this undoped

Page 5: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 121

Fig. 3. Experimental setup for free-space THz electro-optic sampling measurements

GaAs layer is regarded as the channel of a transistor. An ohmic contact wasfabricated by depositing and annealing AuGeNi alloy on the back surface ofthe sample. Then, a semitransparent NiCr Schottky film was deposited on thesurface to apply a DC electric field, F , to the undoped GaAs layer. The roleof the top and bottom contacts is two-fold; first, by using these electrodes, theexternal bias electric field, F , can be applied to the undoped GaAs regions.Secondly, these electrodes provide paths for photoexcited electrons and holesto leave the undoped regions before the next fs light pulse arrives. Thisis quite important in time-domain THz experiments, since if photoexcitedcarriers accumulate in the undoped region, the internal electric field deviatesfrom the planned value due to a field-screening effect, as discussed in Sect. 5.

Femtosecond laser pulses from a mode-locked Ti:Al2O3 laser whose centerwavelength was 790 nm were used to generate electron–hole pairs in the bi-ased undoped region of the diode. The pulse width was approximately 100 fs.The excitation power was set to 30 mW and the illuminated surface area wasapproximately 1 mm × 2 mm. When carriers are accelerated by DC electricfields, they emit EM radiation in the THz regime, whose electric-field compo-nent, ETHz, is proportional to the time derivative of the electron–hole relativevelocity, ∂ve−h(t)/∂t, as schematically shown in Fig. 2b. The free-space THzelectro-optic (EO) sampling technique was used for detecting the THz radi-ation, as illustrated in Fig. 3. The EO sensor used in this experiment was a100 µm thick (110) ZnTe crystal. The spectral bandwidth of this sensor wasapproximately 4 THz (time resolution ≈ 250 fs) [12, 13, 14, 15, 16]. The inter-nal electric field, F , in the undoped depletion region was estimated from thevoltage applied to the m-i-n diode and the surface Schottky barrier height(0.75 eV) determined from an I–V measurement. All the experiments wereperformed at 300 K.

The THz waves emitted from the 1 µm thick GaAs intrinsic region inthe m-i-n diode was measured at various F and its temporal traces are plottedin Fig. 4a [11]. At lower fields, ETHz gradually increases and, then, decreases

Page 6: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

122 Yutaka Kadoya and Kazuhiko Hirakawa

-0.2 0.0 0.2 0.4 0.6 0.8 1.00.0

2.0

4.0

6.0

8.0

47 kV/cm 32 kV/cm 22 kV/cm 17 kV/cm 12 kV/cm 7 kV/cm 4 kV/cm

E THz (

arb.

uni

ts)

Time (ps)

e-h

velo

city

(107 c

m/s

)

(a)

(b)

T = 300 K

Fig. 4. Bias-field depen-dence of (a) the amplitudeof THz electric field, ETHz,emitted from the GaAs m-i-n diode and (b) the transientelectron–hole relative veloc-ity, ve−h(t), obtained by inte-grating ETHz with respect totime [11]

close to zero. At higher fields, ETHz increases more abruptly and its bipolarfeature becomes pronounced. Figure 4b shows the transient velocity obtainedby integrating ETHz with respect to t. When F = 4 kV/cm, ve−h(t) graduallyreaches its steady-state value about 0.3 ps after the excitation. As F is in-creased, the carrier acceleration increases and, furthermore, ve−h(t) starts toexhibit a pronounced overshoot behavior due to intervalley scattering. Sincethe effective mass of electrons is less than that of holes, it has generally beenconsidered that the observed THz signal is dominated by electron transport.We will come back to this point later.

Figure 5 shows the transient velocity of electrons in bulk GaAs calcu-lated by using the EMC method [17, 18]. When compared with Fig. 4b, thequalitative features are similar, indicating that the THz signals are indeeddominated by electron transport. However, when we look at the data moreclosely, a few discrepancies are noticed. First, experimental velocity tracesshow broader overshoot peaks. This is due to the limited detection band-width (≈ 4 THz) of our THz measurement system. Secondly, experimentallyobserved velocities show a gradual decrease after the initial acceleration atlow electric fields (for example, F = 4 kV/cm), while the EMC results donot. This point will be discussed in more detail in the next section.

Page 7: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 123

Fig. 5. Transient electron veloc-ity in bulk GaAs calculated forvarious electric fields by the en-semble Monte Carlo method [11]

2.3 Effect of Sample Geometry on the THz Time-Domain Data

The observed gradual decrease in the transient velocity determined by theTHz measurements shown in Fig. 4b can be explained by the effect of finitesample thickness. Considering the absorption coefficient (α ∼ 104 cm−1) ofGaAs at the photoexcitation wavelength (≈ 790 nm), the 1 µm thick intrinsicregion of the m-i-n diode is almost uniformly illuminated. After the impul-sive photoexcitation, the photoexcited electrons and holes are accelerated bythe electric field and start drifting in opposite directions. However, as theydrift and reach the electrodes, the number of carriers that contribute to theincrease of polarization starts decreasing. Consequently, the apparent carriervelocity obtained from the THz measurements decreases.

In order to quantify such discussions, it is necessary to perform EMCcalculations that take into account the actual sample geometry. In our EMCcalculations, we assumed that the sample surface was illuminated by a fslaser pulse (λ ≈ 790 nm) and electrons and holes with an exponential densityprofile expected from the absorption coefficient were instantaneously createdin the 1 µm thick intrinsic bulk GaAs layer. The calculations were done bycomputing, iteratively, (1) the potential distribution in the biased m-i-n diodeby using a drift-diffusion model and (2) transient velocities of photoexcitedelectrons and holes by using the EMC method. We assumed that carriersthat reach the electrodes vanish and lose the contribution to the ensembleaverage velocity. In this calculation, two valleys (Γ : 0.067m0, Γ : 0.35m0)were assumed for electrons and one valley (Γ : 0.45m0) for holes. The scatter-ing mechanisms included in the calculation were scattering by polar opticalphonons, acoustic phonons, and impurities (1014 cm−3).

Figure 6a schows the comparison between the experimental and EMC re-sults at F = 4 kV/cm. After an initial acceleration, the velocity trace, ve−h(t),obtained from the THz data shows a gradual decrease for t > 300 fs. The cal-culated transient velocities are plotted by a broken line and a dotted line forelectrons and holes, respectively, and the relative velocity, ve−h, is plotted

Page 8: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

124 Yutaka Kadoya and Kazuhiko Hirakawa

0.0 0.2 0.4 0.6 0.8 1.0 1.20.0

2.0

4.0

0.0

2.0

4.0

0.0

2.0

4.0

6.0

Vel

ocity

(107 c

m/s

)

Time (ps)

(c) 12 kV/cm

(b) 7 kV/cm

(a) 4 kV/cm experiment electrons holes electrons + holes

T = 300 K

Fig. 6. Comparison betweenthe transient carrier veloc-ity determined from the THzdata (circles) and the ensem-ble Monte Carlo results at(a) E = 4kV/cm, (b) E =7kV/cm, and (c) E =12kV/cm. The dashed lineand the dotted line denote thetransient velocities for elec-trons and holes, respectively.The solid lines denote thetransient electron–hole rela-tive velocities [11]

by a solid line. As seen in the figure, the calculated ve−h indeed exhibits aslight decrease for t > 400 fs. The agreement between the experimental re-sults and the EMC calculations are reasonable, confirming that the apparentreduction in ve−h(t) originates from the carrier sweep-out effect due to finitesample thickness. This means that even when the channel region is not so thin(≈ 1 µm), the transient velocities obtained by THz experiments are stronglyaffected by the geometrical effect. A similar calculation was done for higherfields (F = 7 kV/cm and 12 kV/cm) and the results are plotted in Figs. 6band c, respectively. The agreement between the THz data and the EMC calcu-lation is indeed excellent, indicating that, if actual experimental conditions(sample geometry, photoexcitation condition, etc.) are properly taken intoaccount, the EMC calculations give a very good description of the transientcarrier velocities determined from THz measurements [11]. The calculationsalso show that the contribution of hole transport is small (only ≈ 10%) forF < 12 kV/cm.

2.4 Field-Dependent Carrier Velocities in Steady States

Now, let us look into the steady-state velocities. In Fig. 7, the steady-statevelocities determined from the THz data in Fig. 4b are plotted, as circles,

Page 9: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 125

0 10 20 30 40 500.0

1.0

2.0

3.0 experiment electrons holes electrons + holes

Ste

ady

stat

e ve

loci

ty (1

07 cm

/s)

Electric field (kV/cm)

Fig. 7. Steady-state veloci-ties determined from Fig. 1bare plotted by circles. Alsoplotted in the figure are thecalculated steady-state veloc-ities for electrons (dashedline), holes (dotted line), andthe sum of the two (solidline) [11]

as a function of F . Because accurate determination of steady-state velocitieswas not possible due the carrier sweep-out effect discussed above, we simplyplotted ve−h(t) at the time when the initial velocity transient (acceleration,overshoot, etc.) and the velocity linearly extrapolated from large t cross. Thesteady-state carrier velocity increases with increasing F and, then, starts todecrease above 7 kV/cm, known as negative differential velocities. In the fig-ure, the dotted line denotes the steady-state velocity of electrons calculatedby the EMC method. Comparing these two, the agreement is reasonable be-low 7 kV/cm. However, the discrepancy becomes significant above 20 kV/cm.In the figure, we also plotted, by a broken line, the calculated steady-statevelocity for holes. The steady-state velocity of holes increases much moregradually than that of electrons because of the larger effective mass, but be-comes comparable when F exceeds 20 kV/cm. The relative velocity given bythe sum of both electron and hole steady-state velocities is plotted by a solidline. As seen in the figure, the agreement with the experimental and EMCresults is very reasonable [11].

3 Bloch Oscillation and THz Gain in SemiconductorSuperlattices

3.1 Bloch Oscillation in Semiconductor Superlattices

In an ideal periodic potential system placed in a uniform static electric field,electrons oscillate both in momentum-space and real-space. This oscillationphenomenon is called “Bloch oscillation (BO)” and corresponds to the Braggreflection of electron wave packets at the edge of the Brillouin zone, whichwas predicted by Bloch [19] and Zener [20]. Bloch oscillating electrons moveback and forth in the energy band at the Bloch frequency ωB (= eFa/;e: elementary charge, : reduced Planck constant, F : applied DC electric

Page 10: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

126 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 8. (a) Superlattice (SL) structure, (b) semiclassical E-k dispersion in unbiasedSLs, and (c) formation of Wannier–Stark ladders in a biased semiconductor SL

field, a: lattice constant). In actual bulk crystals, however, since the energybandwidth is of the order of a few eV, the carriers are scattered far beforereaching the edge of the Brillouin zone and, therefore, BOs cannot be seen.

Owing to recent progress in crystal growth technologies, it has becomepossible to grow periodic stacks of ultrathin semiconducting materials withan atomic-level precision. When two semiconducting materials with differentelectron affinities are periodically stacked, as shown in Fig. 8a, the size of theBrillouin zone is reduced to ±π/d (d: the thickness of one period of the struc-ture; typically ≈ 10 nm) (Fig. 8b). In such structures, called semiconductorsuperlattices (SLs), cosine-like modified energy bands, called “minibands”,with a width of the order of a few tens of meV are formed and, when a DCbias field is applied, electrons are expected to perform BOs, as predicted byEsaki and Tsu [21]. In a quantum-mechanical picture, when an electric fieldis applied to an ideal SL, the miniband is split into equally spaced Wannier–Stark (WS) ladder states with an energy separation of ωB, as schematicallyillustrated in Fig. 8c.

Since the Bloch frequency edF/ can be easily tuned up into the THzrange, the idea of using BOs for realizing ultrahigh frequency oscillators(often called “Bloch oscillators”) has attracted much attention. Since thefirst proposal by Esaki and Tsu [21], considerable effort has been made to

Page 11: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 127

Fig. 9. Superlattice sample structures

search for BOs and obtain THz emission. Ultrafast time-domain experimentson high-quality samples unambiguously demonstrated that electrons in SLminibands perform at least a few cycle Bloch oscillations [6, 22, 23, 24, 25].

However, more recently, it has been recognized that a more fundamentalquestion is whether BO is useful in generating/amplifying EM waves [26].In this section, we would like to discuss this problem and present a strongexperimental support for the THz gain due to Bloch oscillating electrons inSLs. For this purpose, we will show that time-domain THz spectroscopy isvery useful in determining gain/loss spectra in the THz regime [27].

3.2 Superlattice Samples and Time-Domain THz AutocorrelationSpectroscopy

The SL samples studied in this section were GaAs/Al0.3Ga0.7As SL m-i-n(metal-intrinsic-n+-type) diode structures (see Fig. 9). The samples wereprepared by growing 500 nm thick undoped GaAs/Al0.3Ga0.7As SL layerson n+-GaAs substrates by molecular beam epitaxy. We designed the GaAswells (6.4 nm) and Al0.3Ga0.7As barriers (0.56 nm) so as to set the first mini-band width, ∆1, to be 100 meV for sample 1. The other sample (sample 2)had 8.2 nm thick GaAs wells and 0.8 nm thick Al0.3Ga0.7As barriers, and∆1 was designed to be 50 meV. In both samples, the second miniband wasseparated by a 40 meV wide minigap. The top contacts were formed by de-positing a semitransparent 4 nm thick NiCr Schottky film and the bottomohmic contacts were formed by alloying AuGeNi/Au.

When a fs laser pulse excites the sample, electron–hole pairs are opti-cally created in the miniband. Due to an applied electric field, F , the carriers

Page 12: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

128 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 10. (a) The contour plot of the derivative spectra of the interband photocur-rent, ∂Iph/∂(hν), measured on sample 2 as a function of the applied bias voltage.The horizontal arrow denotes the photon energy of the fs laser pulses used forthe time-domain THz measurements. The vertical arrows indicate the anticross-ing points due to Zener tunneling into the second miniband. (b) The integratedintensity of the THz emission is plotted as a function of the bias voltage [27]

start drifting and a THz EM wave that is proportional to the carrier accel-eration is emitted into free space. Since the miniband width for heavy holesis much narrower than that for electrons, heavy holes are almost localized.Furthermore, absorption due to light holes is 1/3 that due to heavy holes.Consequently, the motion of electrons dominates the emitted THz signal.

To verify the formation of the miniband and the WS ladder in our samples,we first took weak-excitation interband photocurrent spectra, Iph. Figure 10ashows the contour plot of the derivative of the interband photocurrent spec-trum, ∂Iph/∂(hν), measured on sample 2 (∆ = 50 meV), where hν is theincident photon energy. For +0.8 V > Vb > +0.5 V, only an excitonic ab-sorption at the bottom of the miniband is observed. However, when Vb isreduced below +0.5 V, the spectrum starts showing rich structures that orig-inate from the formation of the WS ladder. Six ladder levels are clearly seen

Page 13: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 129

in the figure, indicating a good quality sample. Even at T = 300 K (not shownhere), although the structures become weaker and broader, the sign of theWS ladder was still discernible.

For time-resolved THz spectroscopy, an autocorrelation-type interfero-metric set up was used [28], i.e., the power of the THz radiation generatedby a pair of time-correlated fs laser pulses was measured as a function ofthe time interval, τ , between the two laser pulses imposed by a Michelsoninterferometer (Fig. 11). When the laser pulses excite the sample, two almostidentical THz EM waves are emitted consecutively. Experiments were per-formed by using 100 fs laser pulses delivered from a mode-locked Ti:Al2O3

laser. The laser pulses were loosely focused onto the sample surface at apolar angle of 70 C. Under a loosely focused condition, the mechanism forthe THz emission is dominated by acceleration/deceleration of photoexcitedcarriers in the SL miniband. The typical pump power used in the experi-ment was 10 mW to avoid field screening. The generated THz emission wascollimated by off-axis parabolic mirrors and focused onto a Si bolometer op-erated at 4.2 K. The optics and the sample were kept in an enclosure purgedwith nitrogen to minimize absorption by atmospheric moisture. In our quasi-autocorrelation measurements the detected signal is a convolution of the twoTHz electric fields consecutively emitted from the sample.

In Fig. 10b, the intensity of emitted THz radiation is plotted by open cir-cles as a function of applied electric field, F . The THz intensity starts risingat around Vb = +0.65 V, indicating that this voltage is the flatband volt-age, Vb0, in this particular sample. At low fields, the emitted THz intensityincreases with increasing F , which is consistent with an ordinary band trans-port picture. However, when F exceeds 15 kV/cm, the THz intensity startsdecreasing with increasing F , indicating that the dipole moment of oscillat-ing electrons decreases with F . In the WS regime, where edF becomes largerthan the scattering broadening, electrons start to be localized on each well-resolved quantized level in the WS ladder (field-induced Stark localization);the spatial amplitude of each level is given by ∆1/edF and is inversely pro-portional to F . Therefore, the observed reduction of emitted THz intensity isclear evidence for the formation of a WS ladder above a critical electric field.

Figure 12a shows the time-domain autocorrelation traces of the THz emis-sion measured at 10 K for various F . The pump photon energy was set tobe 1.55 eV, which is close to the bottom of the miniband, as indicated by ahorizontal arrow in Fig. 10a [25]. The reason why the bottom of the minibandwas pumped will be discussed in the next paragraph. At very low fields, ahump is seen at around τ = 0. However, as we increase F , a few cyclic oscil-lations start to show up. From the F -dependence of the oscillation frequency,the observed oscillations were identified as Bloch oscillations. The Fourierspectra of the time-domain traces are plotted in Fig. 12b. As seen in thefigure, the frequency of BO shifts to higher frequency as F is increased. How-ever, for F > 17 kV/cm, the spectral shape drastically changes and becomes

Page 14: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

130 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 11. Experimental setup for autocorrelation-type THz time-domain measure-ments

very broad. The behavior in the large reverse bias region will be discussed inSect. 3.4.

Figure 13 shows the THz autocorrelation traces measured at F ∼18 kV/cmfor various pump photon energies, ω, at 300 K [25]. The pump photon en-ergy was varied from 1.442 eV (bottom of the miniband) to 1.498 eV (top ofthe miniband) so as to cover the whole miniband. It is noted from the figurethat clear oscillations are observed only when the electronic states close tothe bottom of the miniband are photoexcited. On the contrary, the oscilla-tions are washed out for ω > 1.462 eV. This behavior can be understoodin the following way; when the SL is pumped at the bottom of the mini-band, the photoexcited electrons occupy the initial momentum states withkz ∼ kxy ∼ 0. In this case, when electrons are accelerated by F , they movealmost in phase from kz ∼ 0. However, when the top of the miniband isphotoexcited, not only electrons with kz ∼ π/d and kxy ∼ 0 but also thosewith small kz and large kxy are photoexcited. Hence, the initial momentum ofelectrons in the acceleration direction, kz , is not uniquely determined. Con-sequently, the phase of BOs is not well defined and clear oscillations are notvisible. Thus, the visibility of BOs is strongly dependent on the preparation

Page 15: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 131

Fig. 12. (a) Autocorrelation traces of the THz emission and (b) their Fourier spec-tra measured on sample 2 for various bias electric fields from 1kV/cm to 29 kV/cmin 2 kV/cm steps. The traces are shifted for clarity [27]

of the initial wave packets by fs laser pulses [25]. It is also noted that a fewcyclic BOs are observed at 300 K even when the miniband width is larger thanthe LO-phonon energy, indicating that BO is a rather robust phenomenon.

Page 16: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

132 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 13. Autocorrelation traces of the THz emission from the SL sample measuredat F = 18 kV/cm for various pump photon energies [25]

3.3 Time-Domain Determination of High-Frequency CarrierConductivities in Superlattices

Owing to the progress in fine epitaxy techniques and fs laser spectroscopy, ul-trafast time-domain experiments on high-quality SL samples unambiguouslydemonstrate that electrons in SL minibands perform at least a few cyclicBloch oscillations [6, 22, 23, 24, 25]. However, the idea of a Bloch oscillatorwas challenged in a more fundamental way by asking whether a BO is usefulin generating/amplifying EM waves [26]. When an electric field is appliedto an ideal SL, the miniband is split into equally spaced WS ladder stateswith an energy separation of ωB, as schematically illustrated in Fig. 8c.Because of the translational symmetry of the structure, the electron popula-tion on each level is identical, too. Since the emission and absorption of EMwaves take place between the adjacent two WS levels, they occur at the samefrequency with the same intensity and, hence, perfectly cancel each other.Consequently, an ideal SL has no net gain or loss for EM waves [26].

However, when scattering exists in the system, the situation changes; inthe presence of scatterers, new transition channels, i.e., scattering-assistedtransitions, become available, as shown in Fig. 14. Since the scattering-assisted emission (absorption) occurs at frequencies slightly lower (higher)than ωB, gain (loss) for EM waves is expected below (above) ωB. This meansthat a Bloch oscillator is not a device that generates an EM wave at ωB, butis a medium that has a broad gain band below ωB [29, 30].

The direct experimental proof of such broad THz gain is to show that thereal part of the carrier conductivity, Re[σ(ω)], becomes negative in the THzrange. The conventional method for determining Re[σ(ω)] in the THz range

Page 17: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 133

Fig. 14. (a) Schematic illustration of the WS ladder in a biased SL with scatterers.(b) Inplane dispersion of WS levels and scattering-assisted emission/absorptionprocesses

is the measurement of free-carrier absorption in doped SLs under bias electricfields by using, for example, Fourier transform spectrometers [31, 32]. If wewant to do so, we immediately face the following two problems; 1. For suchmeasurements, a large sample size, typically a few mm2, is necessary. Then,an extremely large current flows through the sample under biased conditions.2. The profile of internal electric field in doped SLs becomes very nonuniformdue to the formation of high-field domains. These two inevitable problemshave made the conventional measurements practically impossible so far.

We have proposed a new experimental method to determine carrier con-ductivities in the THz region [27]. For determining the high-frequency elec-tron conductivity in SL minibands, we took a totally different approach; weused undoped SL samples in order to avoid a large current density and high-field domain formation, and took all the necessary information within a fewps, by noting that the time-domain THz spectroscopy inherently measuresthe step response of the electron system to the applied bias electric field andthat the Fourier spectra of the THz emission is closely related to σ(ω).

Let us discuss what the THz emission spectra shown in Fig. 12b represent.In the THz-emission experiment, we first set a DC electric field F in the SLand, then, shoot a fs laser pulse at the sample at t = 0 to create a step-function-like carrier density, n0Θ(t), in the sample. Subsequently, THz wavesare emitted from the accelerated photoexcited electrons. Now, by using our

Page 18: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

134 Yutaka Kadoya and Kazuhiko Hirakawa

imagination, we will view the experiment in a different way. Let us performthe following thought experiment; we first set electrons in the conductionminiband under a flatband condition and, at t = 0, suddenly apply a step-function-like bias electric field, FΘ(t). We note that electrons in the thoughtexperiment emit the same THz radiation as in the actual experiment [27].This fact implies that the time-domain THz experiment inherently measuresthe step response of the electron system to the applied electric field.

By noting this important conclusion, we formulate the THz-emissionprocess as follows; the creation of a step-function-like carrier density by fslaser pulses in the actual process is replaced with the application of a step-function-like electric field in the thought experiment as,

F (t) = FΘ(t), F (s) =F

s. (1)

Here, the variables with ˜ denote their Laplace transformation. The transientcurrent by the photoexcited electrons in the miniband is given by

J(s) = σ(s)F (s) . (2)

Then, the emitted THz electric field, ETHz, can be obtained as,

ETHz ∝∂J(t)

∂t=

12πi

∫ c+j∞

c−j∞sσ(s)ets F

s= σ(t)F , (3)

where σ(t) is the electron conductivity in the time domain. Equation (3) tellsus that since the spectra shown in Fig. 12b are the Fourier spectra of theTHz autocorrelation traces they are proportional to |σ(ω)|2F 2.

In the following, we will compare the observed THz spectra with the con-ductivity theoretically predicted for the Bloch oscillating electrons by Ktitrovet al. [33] by solving the semiclassical Boltzmann transport equation

σ(ω) = σ01 − ω2

Bτmτe − iωτe

(ω2B − ω2) τmτe + 1 − iω (τm + τe)

(4)

with

σ0 =σ00

1 + ω2Bτmτe

,

where τm and τe are the momentum and energy relaxation times of elec-trons, respectively, and σ00 is a constant. The top panels of Fig. 15 showthe calculated real and imaginary parts of the miniband conductivity, andthe bottom panels show the measured THz spectra and calculated |σ(ω)|2for two representative bias conditions. τm and τe in the calculation were de-termined as 0.15 ps and 1.5 ps, respectively, from numerical fitting. ωB atF = 1.0 kV/cm was simply set to be edF/. However, we found that thepeak position of the THz spectrum at 11.1 kV/cm is slightly lower than the

Page 19: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 135

Fig. 15. The top panels show the calculated real (solid line) and imaginary (dashedline) parts of σ(ω) for the SL miniband. The bottom panels show the calculated|σ(ω)|2-spectra (dashed line) and the THz emission spectra (dots) measured forsample 2 by time-domain THz spectroscopy; (a) F = 1.0 kV/cm and (b) F =11 kV/cm [27]

value simply expected from edF/ (2.41 THz), which is due to anticrossingwith the WS ladder states of the second miniband, as will be discussed later.We set ωB at F = 11.1 kV/cm to be 1.96 THz, which gives the best fit to theexperimental data.

At very low field [F = 1.0 kV/cm, ωB(τmτe)0.5 = 0.66], the conductivityspectra and the THz emission show more or less a Drude-like behavior (seeFig. 15a). However, as F is increased, (4) predicts that a negative Re[σ(ω)] be-low ωB develops and that a clear dip in Im[σ(ω)] appears at ωB. Consequently,the expected |σ(ω)|2 spectrum becomes asymmetric with an enhanced low-frequency spectral component. It should be noted in Fig. 15b that the THzemission at F = 11.1 kV/cm [ωB(τmτe)0.5 = 5.8] indeed has a characteris-tic asymmetric spectral shape and that the agreement between the observedTHz emission and the calculated |σ(ω)|2 is excellent, indicating that ETHz

is indeed proportional to σ(ω) of Bloch oscillating electrons. Furthermore,the calculated Re[σ(ω)] strongly suggests that the THz gain persists at leastup to 1.7 THz [27]. Similar experiments were performed for another sample(∆1 = 100 meV). Figure 16 summarizes the THz spectra and the calculated|σ(ω)|2 for various F . Except for the case at F = 2.6 kV/cm, where the THz

Page 20: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

136 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 16. The calculated |σ(ω)|2 spectra (dashed line) and the THz emission spectra(dots) measured for sample 1 are shown for four bias electric fields; (a) 2.6 kV/cm,(b) 6.5 kV/cm, (c) 10.4 kV/cm, and (d) 14.4 kV/cm. The sharp dips at 0.5 THz inthe measured data are due to water-vapor absorption [27]

emission is quite weak and the signal/noise ratio is low, the excellent agree-ment between experiment and theory further confirms the above conclusion.

3.4 Zener Tunneling into Higher Minibandsand High-Frequency Limit of the Bloch Gain

Now, we would like to discuss what limits the maximum frequencies of theBloch gain in SLs. The good agreement between theory and experiment dis-cussed above was obtained only up to 13 kV/cm for the THz data shownin Fig. 12b. As F is further increased, the period of the Bloch oscillationbecomes shorter but, at the same time, the amplitude gradually diminishes.Concomitantly, the Fourier spectra become very broad (see Fig. 12b). We at-tribute this behavior to the anticrossing of the WS states of the first minibandwith those of the second miniband, i.e., Zener tunneling into the second mini-band. Let us take a look at Fig. 10b again. It is noted that the THz emissionbecomes brighter with increasing field for Vb > −0.8 V(F > 29 kV/cm). Thiscontradicts what is expected from simple field-induced Stark localization in

Page 21: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 137

a single miniband and indicates that electrons in the first miniband find anew conduction channel. This new channel is the Zener tunneling into thesecond miniband. Further support can be seen in the interband photocurrentspectra shown in Fig. 10a, where an anticrossing feature can be observed inthe high-bias region, as indicated by vertical arrows [34]. From these data,it is concluded that the Zener tunneling into the higher miniband sets thehigh-frequency limit to the Bloch gain.

Hence, it is of prime importance to clarify the electronic structures ofhigher-lying minibands/minigaps in SLs and understand the electron trans-port in a strong bias field range. In this section, we will discuss the Zenertunneling effect in biased wide-miniband GaAs/Al0.3Ga0.7As SLs in moredetail.

The samples were the same GaAs/Al0.3Ga0.7As SL m-i-n diode structuresas those used in the previous section. We simply measured the intensity ofthe THz radiation by using a Si bolometer. Experiments were performed byusing 100 fs laser pulses delivered from a mode-locked Ti:Al2O3 laser. Thelaser pulses were loosely focused (φ = 1 mm) onto the sample surface at apolar angle of 70 C. The pump photon energy was set to excite carriers atthe bottom of the first miniband. The generated THz emission was collimatedby off-axis parabolic mirrors and focused onto a Si bolometer.

Figure 17a shows the bolometer output for the THz radiation emittedfrom sample 1 as a function of the bias voltage, Vb, applied to the SL. TheTHz intensity starts rising at around Vb = +0.83 V, indicating that thisvoltage is the flatband voltage, Vb0, in this particular sample. The appliedelectric field F was estimated by subtracting Vb0 from Vb and dividing bythe total thickness of the SL region of 500 nm.

Two distinct regions can be seen in the figure; one region (I) is the THzemission for 0 kV/cm to 30 kV/cm. The THz emission in region I is due tointraminiband electron transport [25]. When F 8 kV/cm, strong interwellcoupling leads to the formation of energy minibands. Furthermore, the Blochenergy (≡ eFd; d: the SL period) is less than the scattering broadening ofthe WS levels. Therefore, electrons drift in a quasicontinuous miniband. Froma closer look at Fig. 17a, a weak kink in the THz emission curve is noticedat ≈ 8 kV/cm. For F 8 kV/cm, the miniband splits into equally spaced WSstates and the wave functions are gradually localized (WS localization). Inthis intermediate field region, clear Bloch oscillations are observed in the time-domain THz traces [25, 27]. When F exceeds ≈ 18 kV/cm, the THz intensitystarts rolling off due to stronger WS localization. The other region (II) isa broad THz-emission band observed for 30 kV/cm to 110 kV/cm. When Fexceeds 30 kV/cm, the reduction of THz intensity saturates and the emissionbecomes brighter again. Furthermore, quasiperiodic structures are observedon top of the broad peak [35]. The dashed curve is a blowup of the THzintensity in the high field region. Since the THz intensity is proportional tothe carrier acceleration, the THz emission becomes brighter whenever carriers

Page 22: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

138 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 17. (a) The bias-field dependence of the THz intensity measured for sample 1.The dashed curve shows a blowup of the THz intensity in the high-field region.The arrow at ≈ 8 kV/cm shows the position of a weak kink in the THz-intensityspectrum. (b) The inverse of the resonance electric fields plotted as a function ofthe index number (see Fig. 18) [35]

find new transport channels. We attribute the broad feature and the finestructure to the nonresonant and resonant interminiband Zener tunnelingeffects, respectively [34, 36, 37]. In the following, we will focus our attentionon the fine structure in the THz-emission traces.

Let us consider the band diagram under a strong bias field, where theinterminiband Zener tunneling takes place. Figure 18 shows a schematic il-lustration of the SL band diagram in a high-field condition. Equally spacedWS levels with an energy separation of eFd are formed for both the first andthe second minibands. It is also noted that the energy separation between

Page 23: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 139

Fig. 18. Schematic illustration of the conduction-band diagram of the SL under astrongly biased condition [35]

the first and the second quantized energy levels in each quantum well (QW),E21 (= E2 − E1), is almost field independent, since the quantum-confinedStark effect is weak in narrow QWs. Consequently, the energy level of theground state E1 in the 0-th well is aligned with the excited state E2 inthe n-th well when F satisfies the condition; E21 = neFd, where the index ndenotes the position of the QW numbered with respect to the QW wherethe wave function of the ground state is localized (see Fig. 18). Under sucha condition, electrons in the first miniband can resonantly tunnel into theexcited state in the n-th well and the THz emission is enhanced. In Fig. 17b,the inverse of the resonance electric field, Fr, indicated by arrows in Fig. 17ais plotted as a function of n. The linear dependence between F−1

r and theindex n indicates that the quasiperiodic peak structure is indeed due to theinterminiband resonant Zener tunneling between the WS ladders associatedwith the first and the second minibands. Resonant Zener tunneling up intothe 8th-nearest neighbor QW was clearly resolved, indicating that the wavefunction in the SL coherently extends over ≈ 60 nm (8 periods × 7 nm) [35].

4 THz Radiation From Semiconductor Microcavitiesin Strong Exciton–Photon Coupling Regime

4.1 Cavity-Polaritons and the Idea of THz Wave Radiation

When quantum wells (QWs) are embedded in a planar wavelength-scale op-tical cavity (microcavity), the light–matter interactions are largely modified

Page 24: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

140 Yutaka Kadoya and Kazuhiko Hirakawa

owing to the confinement of the light field (a review is given, for example,in [38]). In particular, in high-Q semiconductor microcavities (SMCs) at lowtemperatures, where the damping of the cavity-mode photons and the exci-ton coherence are sufficiently slow, coherent coupling between the excitonsand photons brings about the formation of two-dimensional exciton–polaritonstates called cavity-polaritons [39]. The coupling is often called “strong cou-pling”. In such a system, neither the exciton nor the cavity-mode photon,but the two polariton states become the normal modes of the system. Be-cause of the two-dimensional nature of the system, where the motions in thedirection normal to the QW and cavity mirror planes are quantized, the cav-ity and exciton states, from which the polaritons are formed, are uniquelydetermined once an inplane wavevector K‖ is given. Experimentally, K‖ isdetermined by selecting the angle θin of the incident or emitted light withrespect to the normal of the layer plane. Hence, the cavity-polariton modesare observed spectrally as the splitting (anticrossing) in the reflection spectrataken at various θin [39], and the dispersion curve can be drawn directly, asillustrated in Fig. 19.

The angular frequency of the two (upper and lower) polariton modes, ω+

and ω− are expressed as, neglecting the damping terms,

ω± =ωex + ωc

√(ωex − ωc)2 + Ω2

0

2, (5)

where ωex and ωc, both functions of K‖, are the energy of the exciton andthe cavity mode, respectively. Practically, the cavity mode depends stronglyon K‖, whereas the exciton energy is effectively dispersionless for K‖, asillustrated in Fig. 19. The splitting Ω0 at zero detuning, δ = ωex − ωc = 0,which is sometimes called the “Rabi” frequency, is determined by the couplingstrength between the exciton and the cavity mode, and is proportional to thesquare-root of the exciton oscillator strength,

√fosc. In the case of GaAs-

based SMCs, the splitting energy Ω0 is of the order of meV.An excitation of such a system by a short laser pulse can trigger a comple-

mentary oscillation at the frequency Ω/2π = (ω+−ω−)/2π, in the exciton andphoton populations. The oscillation, called normal mode oscillation (NMO),corresponds to the intuitive time-domain picture of polaritons, where the en-ergy is exchanged periodically between the excitons and photons. In a sensethat neither the exciton nor the photon state is the eigenstate of the system,the occurrence of NMO is natural, because the laser pulse is a photonic state.In GaAs-based SMCs, the frequency of NMO, Ω/2π, lies in the THz regime.Alternatively, NMO can be understood as the beating in the polarization orthe light field carried by the two polariton states, at ω+ and ω−. For example,at low excitation, the exciton density can be expressed as the square of thepolarization, namely nex = |Pex|2 = |P (ω+) + P (ω−)|2, which oscillates atthe frequency Ω = ω+ − ω−. Hence, one can note that, to initiate NMO, itis necessary to excite both the polariton modes coherently. NMO has been

Page 25: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 141

Fig. 19. Illustration of the dispersion curve of cavity-polariton modes and thesurface reflection spectra for the incidence angle that corresponds to the inplanewave vector K‖ producing the exciton-cavity resonance

observed in resonant luminescence [40, 41, 42] or pump-probe type measure-ments including four-wave-mixing [40, 43, 44, 45, 46]. In these experiments,however, all the relevant light beams are in the frequency range that is con-fined in the microcavity, and the signal could suffer from the higher-orderinterference effect.

On the other hand, it has been pointed out that THz EM waves areradiated from cavity-polaritons undergoing NMO [47]. The idea is as follows.In the presence of a static electric filed in the QW, the electron and holewavefunctions are displaced in the opposite direction, as illustrated in Fig. 20.Since such a polarized exciton carries a static dipole, µex = e(zh − ze), wherezh = 〈h|z|h〉 and ze = 〈e|z|e〉 are the average position of a hole and electron,respectively, the oscillation in the exciton population Nex due to NMO resultsin an oscillation of macroscopic dipole D ∝ Nexµex, which is expected toradiate EM waves, according to the well-known electromagnetic theory, at thefrequency Ω/2π ∼ THz. Therefore the radiation of the THz waves expressedas

ETHz ∝∂2

∂t2[Nexµex] =

∂2Nex

∂t2e(zh − ze) , (6)

can manifest the oscillation of Nex associated with the NMO. In contrastto the pump-probe-type measurements mentioned above, the THz-emissionexperiments, where the signals are not confined in the cavity, allow us toobserve the signal free from the unwanted interference effects.

The radiation of THz waves from cavity-polaritons can be also understoodas a kind of optical rectification in the second-order nonlinear process, since itcomes from the oscillation nex, which is proportional to |Pex|2 as mentionedabove. From the viewpoint of optical nonlinearity, to obtain second-orderpolarization, spatial inversion symmetry of the system must be broken, whichis indeed the case (∆z = 0) in the QW excitons under a static electric field,as shown in Fig. 20.

Page 26: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

142 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 20. The mechanismof THz wave radiation fromcavity-polaritons undergoingNMO

4.2 The Microcavity Samples and the Cavity-Polariton Modesunder a Static Electric Field

The samples used here are the SMC devices [48,49], as illustrated in Fig. 21a,where two GaAs QWs separated by an AlAs barrier are embedded nearthe center of the half-wavelength cavity made of highly reflective distrib-uted Bragg reflectors (DBRs), which is composed of alternating quarter-wavelength AlAs and Al0.2Ga0.8As layers. The number of layers in the air-and substrate-side DBR are 33 and 58, respectively, for which the Q-valueof the empty cavity was evaluated to be ≈ 3700 from the surface reflectionspectrum of the sample in which the QW can be regarded to be transpar-ent (ωc ωex). The substrate-side DBR is doped with Be while in the air-sideDBR, only a part (12 nm) of an AlGaAs layer is doped with Si, to reduce theabsorption for the THz wave.

A DC voltage, VDC was applied between the Si-doped conducting Al-GaAs layer, and the p-type substrate. In Fig. 21b, the energy levels of theexciton-cavity coupled system is schematically illustrated as functions of theDC electric field in the QW, EQW. Here, only a lowest heavy-hole excitonstate is considered. With the increase of EQW, the exciton transition fre-quency ωex is redshifted through the quantum-confined Stark effect (QCSE),while the cavity resonance ωc is determined by the incidence angle of the exci-tation light θin, independently of EQW. Around the exciton-cavity resonanceωex ∼ ωc, the anticrossing shows up, corresponding to that in the polaritondispersion curve as shown in Fig. 19. By changing EQW (Vdc), while keepingωc (θin), the polariton modes with various detuning δ = ωex − ωc are real-ized. On the other hand, by the combined change of Vdc and θin, δ = 0 isachieved at various values of EQW. In the tuning of EQW, fosc varies also dueto QCSE so that Ω0 can be also modified. Hence, in the experiment, Ω wasvaried through the change of δ and Ω0.

Page 27: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 143

Fig. 21. (a) Schematic drawing of the high-Q SMC device used for THz waveradiation. (b) Schematic drawing of the polariton modes under a static electricfield

Fig. 22. (a) Examples of the detected THz signal for three cases of the modesplitting, all at δ = 0 realized by the combined tuning of EQW and θin. (b) Reflectionspectra of the excitation laser pulses, for the three cases shown in (a). The two dipsin each spectrum correspond to the polariton modes

4.3 Observation of THz Wave Emissions from Cavity-Polaritons

In the THz-wave-generation experiment, the sample was cooled to 10 K in acryostat, and the time-domain traces of the THz waves were measured in astandard THz time-domain spectroscopy (TDS) setup using a mode-lockedTi-sapphire laser and a photoconductive antenna.

Page 28: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

144 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 23. Frequency of the THzwave is plotted as a function ofthe mode splitting ΩR observedin the reflection spectra. The solidline shows the frequency ΩR/π.The dashed curve shows the fre-quency Ω/π, where the internalvalue of the mode splitting Ω wascalculated from ΩR, taking intoaccount the damping effect

Figure 22a shows the examples of the detected signals for three cases ofthe mode splitting at the exciton-cavity resonance, δ = 0, realized by thecombined tuning of Vdc and θin, for which the splitting depends on EQW

through the change of Ω0. The values of ΩR indicated in the figure are thefrequency difference between the two polariton modes seen in the reflectionspectra as shown in Fig. 22b. The ΩR value is very close to, but can beslightly different from the true mode splitting Ω = ω+ − ω− (= Ω0 forδ = 0) calculated from (5) as mentioned below. In all the cases, several-cycleoscillations are clearly seen, manifesting the oscillation of exciton populationassociated with the NMO. The dashed lines connecting the correspondingmaxima and minima were drawn to guide the reader’s eye. Obviously, forthe smaller splitting ΩR(∼ Ω), the oscillation gets slower, as expected forthe NMO. Larger damping of the oscillation at higher EQW corresponds tothe shallower polariton-mode dips seen in Fig. 22b. In the present case, theincrease of the damping rate is due to the increase of the dephasing rate ofthe exciton.

The frequency of the THz signals evaluated as the inverse of the oscil-lation period is summarized in Fig. 23 as a function of ΩR. As mentionedabove, Ω (and ΩR) can be also modified with the change of δ. The datapoints in the figure include both the cases of exciton-cavity resonance (δ = 0,Ω0: varied) and offresonance (δ: varied, Ω0 ∼ constant). For the offresonantcases, the exciton was detuned to the higher-energy side of the cavity mode(ωc < ωex), to avoid the influence by the light-hole exciton and continuumstates. Basically, the experimental data points are fitted quite well by thesolid line representing the values of ΩR/2π.

The observed frequency, however, deviates from the straight line, particu-larly in small ΩR cases. Recalling that the decrease of ΩR is realized at δ = 0by the decrease of Ω0 with the increase of EQW (see Fig. 22), we can drawan important conclusion. In general, ΩR is different from Ω, which is the in-trinsic frequency of NMO, because of the finite linewidths (dampings) of the

Page 29: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 145

cavity and exciton modes, particularly when the linewidth is not sufficientlysmaller than Ω [50]. In the present sample, the linewidths of the excitons andthe cavity resonance are about 2 meV and 0.6 meV, respectively, for whichthe arithmetic average, 1.3 meV, is not very small compared with Ω ≈ 4 meV.The exciton linewidth increases gradually with the increase of EQW [47]. Tak-ing into account the linewidth, we calculated Ω for each ΩR and plotted it bythe broken curve in Fig. 23. Though the fitting is not perfect, the differencebetween ΩR and Ω accounts qualitatively for the observed deviation of thedata points from the ΩR line. In other words, the frequency of the emittedTHz wave is not that of the beating between the optical fields, whose fre-quency is given by ΩR, but is equal to the frequency of the NMO. Hence, wecan say that, in the THz-emission technique, the oscillation of the excitonpopulation due to NMO is directly observed.

In order to confirm the prediction, (6), the EQW and excitation powerdependencies of the EM wave amplitude were studied. In the experiment,the exciton-cavity resonance (δ = 0) was kept constant by the tuning of theincidence angle, or EQW, and the center frequency of the excitation pulseswas adjusted to the polariton doublets. Figure 24a shows the EM wave ampli-tude as a function of EQW. The amplitude increases monotonically with theincrease of EQW, demonstrating the significance of the application of a staticelectric field. In the figure, the relative change of the displacement betweenthe electrons and holes ∆z = 〈1hh|z|1hh〉−〈1e|z|1e〉, which is proportional tothe exciton self-dipole µex, calculated for the 12 nm thick GaAs is also shownby the solid curve. The observed dependence of the THz amplitude on EQW

corresponds quite well to the increase of ∆z, supporting the prediction, (6).Shown in Fig. 24b is the dependence of the amplitude on the excitation laserpower Pex, to which the exciton population Nex is proportional. The datapoints clearly fall on the line ETHz ∝ Pex manifesting again the validityof (6). Then, combining the two dependencies, we can summarize that theTHz waves are radiated as a result of the second-order nonlinear responsecaused by the field-induced break of the inversion symmetry.

4.4 Transition from Strong- to Weak-Coupling Regime

As demonstrated clearly in the preceding sections, the THz waves are radi-ated from the coherent nutation of the exciton population associated withthe NMO of the cavity-polariton system. Such coherent nutation decays dueto the dephasing of the exciton coherence or the leakage of the light fieldfrom the cavity. The effect of the increase in the exciton dephasing has beenshown, though not clearly, in Fig. 22, as the faster damping of the oscillationfor higher EQW. Here, we discuss the influence of the cavity-mode damping.For this purpose, we prepared SMC samples where the number of layers inthe air-side DBR, Nair was reduced to 23 and 13 (originally 33), for whichthe Q-values of the empty cavities are about 1100 and 240, respectively (orig-inally 3700). In the lowest Q-value sample, an anticrossing is not observed

Page 30: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

146 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 24. Amplitude of the THz wave is plotted as a function of (a) the staticelectric field in the QW, EQW, and (b) the excitation power Pex. The experimentswere performed keeping the exciton-cavity resonance (δ = 0). In (a) the calculatedvariation of the relative displacement ∆z between the electron and hole is shownby the solid curve and in (b) the solid line is drawn to show the dependenceETHz ∝ Pex

Fig. 25. THz signal obtained withvarious Q-value samples. The num-ber n in the labels 〈58-n〉 indicatesthe number of layers in the air-sideDBR. The signals were normalized bythe first half-cycle

in the reflection spectrum (not shown). Shown in Fig. 25 are the observedTHz signals for these samples. To see clearly the decay of the oscillation, thesignals were normalized by the amplitude of the first half-cycle. Obviouslythe decay is fast in low-Q samples, and nearly one cycle is left in the lowestQ-value sample. The first one cycle results from the instantaneous creationof the excitons by the laser pulses, which is observed even in the excitationof a single exciton or continuum states in bare QWs [51]. In other words, it is

Page 31: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 147

the oscillation following the first cycle that manifests the coherent nutationof the exciton population.

5 THz Radiation from Bulk SemiconductorMicrocavities

5.1 THz Wave Radiation from Semiconductor Surfaces: A Revisit

The excitation of unbiased semiconductor surfaces by short laser pulses is anattractive method for obtaining THz EM waves [52], because of the simplicityof the setup and the robustness against the excitation-induced damage ofthe emitter. In the case of rather low excitation density, where the opticalrectification effect [53] is negligible, the source of the THz wave is the surgecurrent carried by impulsively excited carriers near the surface [52]. In wide-gap semiconductors such as GaAs and InP where the built-in field is ratherhigh, the current stems from the drift motion of the photoexcited carriers bythe built-in surface electric field. Depending on the excitation pulse widthwith respect to the period of carrier acceleration, the time derivative of thecurrent, to which the far-field THz wave is proportional, is dominated eitherby the variation of the velocity [9] or of the density [52] of the photoexcitedcarriers. In both cases, the strength of the static electric field and the totalnumber of carriers excited in the field region are the important parametersdetermining the amplitude of the THz waves.

The built-in surface electric field is caused by the difference in the Fermilevels between the surface and the bulk region. In most III–V semiconductors,the Fermi level at the surface is pinned at a certain energetic position, while itlies around the conduction band edge in n-doped bulk region. In off-the-shelfbulk semiconductors, only the built-in potential difference can be chosen. Fora given built-in potential difference, a thinner field region is beneficial forobtaining a higher field, while it is unfavorable for exciting larger number ofcarriers, as illustrated in Figs. 26a and b. Hence, there exists a sort of tradeoffrelation between the field strength and the photocarrier number.

In an optical cavity, in contrast, the excitation pulses can be absorbedefficiently during the round trips, even though a thin absorption layer isinvolved, allowing us to realize a rather high static field with only the built-in potential, as illustrated in Fig. 26c. As a result, one may expect largerTHz wave amplitude for a given excitation light intensity. In this section,we show that the microcavity structures, which have been described in theprevious section, can be used also for the enhancement of the THz waves, byreplacing the QWs by a bulk GaAs [54]. In addition, we discuss an influenceof the carrier accumulation on the radiation of THz waves that has not beenconsidered yet.

Page 32: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

148 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 26. Illustration of the potential profile of n-GaAs and SMC near the surfaceat the thermal equilibrium, displaying that (a) the light absorption is efficient butthe field is low in a thick depletion region, (b) the field is high but the absorption isinefficient in a thin depletion region, (c) both a high field and efficient absorptionis realized in the SMC

5.2 Enhancement of THz Wave Generation Efficiency

The cross-sectional structure and the energy-band diagram around the THz-wave generating region of the SMC samples considered here are schematicallydrawn in Fig. 27a. A 182.6 nm thick undoped GaAs and a 50 nm thick Si-doped (∼ 2×1018 cm−3) GaAs are sandwiched by a pair of DBRs, forming aone-wavelength cavity. The substrate-side DBR is doped with Si, while theair-side one is undoped. The THz waves are radiated from the surge currentcreated by the laser pulses in the undoped GaAs layer, where a static electricfield is built-in, as in the case of bare n-type GaAs substrate. In order toobtain a higher field in the undoped GaAs region, the Fermi-level pinningis desired to take place not at the surface but at the interface between theundoped GaAs layer and the air-side DBR. Such an intentional pinning wasintroduced in the sample preparation by interrupting the molecular beamepitaxy growth and exposing the GaAs surface to the room air.

Assuming that the energetic position of the Fermi level at the interface isthe same as that at the bare GaAs surface (≈ 0.7 eV below the conduction-band edge), and that in the n-GaAs region is located at the band edge, thebuilt-in electric field EB at thermal equilibrium is expected to be 38 kV/cm.For the present purpose, not a very high but a low-Q cavity is better, asshown later. Hence, in the experiments, two types of cavity were tested; onehaving 11.5 and 4 AlAs/Al0.2Ga0.8As pairs in the substrate- and air-sideDBRs, respectively, referred to as a (23-8) cavity, and the other having 11.5and 2 pairs in each side, referred to as a (23-4) cavity. The results werecompared with those obtained with a reference sample, whose structure isshown in Fig. 27b, and an InAs sample, in which an undoped 500 nm thickInAs has been grown on semi-insulating GaAs substrate by molecular beamepitaxy. Figure 28 shows the surface reflection spectra of the cavity samplesfor an incidence angle of 45. The cavity resonance is clearly seen around820 nm, where the reflectivity is very low, as expected. In particular, in a(23-4) cavity, nearly zero reflectance is realized.

The THz-wave generation experiments were performed with a standardTHz TDS setup where a mode-locked Ti-sapphire laser and a photocon-

Page 33: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 149

Fig. 27. Schematic drawing of the bulk SMC and reference samples together withthe potential profile at the thermal equilibrium

Fig. 28. Surface reflection spectra of bulk SMC devices for 45 incidence; (a) (23-8)cavity, (b) (23-4) cavity. The solid and dotted curves are the experimental resultsand theoretical predictions, respectively

ductive antenna were used. For the excitation, the laser pulses centeredat ≈ 820 nm were incident on the samples with an angle of 45 for whichthe cavities satisfy the resonant condition. The power reflectance of a (23-8)cavity, a (23-4) cavity, the reference, and InAs samples for this conditionwere 41%, 17%, 40%, and 44%, respectively. The amplitude of the THz wavewas evaluated from the difference between the positive and negative peaks inthe time-domain signals.

Shown in Fig. 29 are the amplitudes of the THz waves obtained with(23-8) and (23-4) cavities, the reference, and InAs samples as functions ofthe excitation power. The excitation spot is about 0.8 mm in diameter. In

Page 34: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

150 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 29. The amplitude of the THzwaves obtained with a (23-8) cavity,a (23-4) cavity, a reference, and InAssamples as functions of the excitationpower [54]

the low-excitation-power regime, the amplitudes of the cavity samples areobviously larger than that obtained with the reference and InAs samples.With the increase of the excitation intensity, the amplitude obtained withInAs increases linearly, while the increase is obviously sublinear in the case ofcavities, even in the low-excitation regime, and the deviation from the linearrelation becomes larger in the high-excitation regime. The sublinear relationcan be explained by the screening of the static field by the photoexcitedcarriers as discussed in the next section. By taking into account the differenceof the reflectance between the samples, and comparing the results at the samenet excitation power, which penetrate into the samples, one can find that thedifference between the cavities becomes negligible and that the amplitudeof the cavities are 2.6 times larger than that of the reference sample, and3.1 times larger than that of InAs.

Assuming the absorption coefficient in GaAs to be α = 1.2×104 cm−1 [55],the absorbed power in the undoped GaAs region in the reference sample isevaluated to be 20% of the net excitation. In the cavity samples, assumingthat all the net power is absorbed inside the cavity (no transmission) andtaking the standing-wave pattern into account, the absorption in the undopedGaAs region can be evaluated roughly to be 75%. From these values, onemay expect an enhancement factor of 3.8, if the amplitude of the THz waveis proportional to the total number of the carriers excited in the undopedGaAs region. The smaller experimental value may be due to the screeningeffect, which exists even in the lowest-excitation case, as indicated by thesublinear dependence of the amplitude on the excitation power.

One point that should be mentioned is that the THz amplitude obtainedwith the reference sample is the same as or even stronger than that withInAs. On the other hand, the THz amplitude obtained with n-type GaAsdoped with Si to 2×1018 cm−3 is about a factor of three smaller than theundoped InAs (not shown here). Hence, the results indicate clearly that the

Page 35: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 151

Fig. 30. The amplitude of THz wavesobtained with various excitation spotsize, as functions of the excitationpower. The dashed line indicates thelinear relationship [54]

THz intensity depends on the doping profile, and that a comparison betweendifferent materials needs care due to the difference of the potential as well asthe electronic structures.

From an application point of view, it is important to realize the enhance-ment even in the high-excitation-power regime. If the photocarrier-inducedscreening is the origin of the sublinear dependence of the THz wave ampli-tude on the excitation intensity shown in Fig. 29, one simple way to increasethe efficiency at a high excitation power is to use a larger excitation spot.Figure 30 shows the excitation-power dependence of the THz amplitude forvarious spot sizes, obtained with a (23-8) cavity, in comparison with the caseof InAs. As anticipated, the deviation from the linear relationship betweenthe excitation power and the amplitude is smaller for a larger excitation spot.In particular, for the 8 mm spot case, the amplitude of the THz wave obtainedwith the cavity increases almost linearly with the excitation intensity, and isstill a factor of 3 (factor of 9 in intensity) larger than that with InAs, at theexcitation of 65 mW. However, as can be seen by a comparison with the lin-ear relationship indicated by the broken line in the figure, the dependence ofthe amplitude on the excitation intensity is still slightly sublinear. Hence, toutilize the cavity-induced enhancement even at higher excitation intensities,a larger spot should be used.

5.3 Field Screening by Carrier Accumulation

In the 8 mm spot case shown in Fig. 30, even at the highest excitation, 65 mW,the carrier density created by a single pulse is about 8×109 cm−2, for whichthe screening field, ≈ 1.2 kV/cm, is much smaller than the equilibrium field,EB ≈ 38 kV/cm. Hence, the screening associated with each excitation pulsecannot explain the sublinear dependence. Rather, for such a case, one mayneed to take into account the accumulation of the photoexcited carriers. Inthe microcavity samples, the photoexcited electrons and holes accumulate

Page 36: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

152 Yutaka Kadoya and Kazuhiko Hirakawa

Fig. 31. (a) The calculation model for the recovery process. (b) The excess carrierdensity, n0, just before the arrival of each pulse (dotted curve) and that, n0 + nP,just after the arrival (solid curve), and the effective electric field Eeff for the THzwave radiation as functions of the carrier density nP created by each pulse [54]

in the n-doped GaAs region and the interface between the undoped GaAsand the air-side DBR, respectively. They may recombine inside the device,through the transportation across the undoped GaAs region. However, if therecombination does not complete between the pulses, an amount of the excesscarriers may be left.

Here, to make the analysis simple, we treat the transportation as thethermionic emissions of electrons from the n-GaAs layer to the DBR/GaAsinterface, considering the simplified potential profile shown in Fig. 31a. Thenonequilibrium voltage ∆V induced by the excess carriers of sheet den-sity ns(t) is expressed as ∆V (t) = [ens(t)/ε]di, where e, ε and di are theelementary charge, dielectric constant, and the thickness of the undopedGaAs layer, respectively. The thermionic current density is expressed as [56]Jte(t) = Js[exp(e∆V (t)/kBT )−1], where Js and kBT/e are the saturation cur-rent density and the thermal energy, respectively. Neglecting the carrier trans-port during each excitation pulse, the time dependence of ns after an impulseexcitation can be obtained by solving the equation, ∂ns(t)/∂t = −Jte(t)/e.Then, for the steady-state situation where the carriers of density nP arecreated periodically by each pulse with the interval, τML = 13 ns, of themode-locking laser, the carrier density n0 remaining just before the arrivalof each pulse is obtained. Using thus-calculated values of n0, the effectivestatic electric field, by which the photoexcited carriers are accelerated, wasevaluated as Eeff = EB − e(n0 + nP/2)/ε, where a half of nP is included as arough estimation of the average field during an excitation pulse.

Figure 31b shows the estimated values of n0, n0 + nP, and Eeff as func-tions of nP, using the values Js = 1.5×10−6 A/cm2 [56], T = 300 K, and thedielectric constant of GaAs, 13.1. One can notice that, even for a weak ex-

Page 37: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 153

citation, nP 1011 cm−2, the accumulated excess carrier density n0 is quitehigh, n0 ∼ 1011 cm−2, and the effective field Eeff is apparently smaller thanthe equilibrium value, 38 kV/cm. This can be understood as follows. In theearly stage of the recovery process, the thermionic current is high because oflarge ∆V . However, with the decrease of ∆V , the current decreases exponen-tially and the recovery process slows down drastically, remaining rather highn0 at the arrival of the next excitation pulse. Hence, the value of ∆V , andhence the current density, at a time long after each excitation pulse, is notdominated by the initial carrier density, n0 and ∆V depend weakly on thecarrier density per pulse, nP. The gradual decrease of Eeff with increasing nP

results in the weakly sublinear dependence of the THz amplitude on the ex-citation power. As indicated by the arrows in the figure, the experiment withthe 8 mm spot and that with the 0.8 mm spot under the weak excitation inFig. 30 correspond to such a regime. Therefore, in the low-excitation-densityregime, the screening of the built-in surface electric field by the accumulatedphotocarriers explains qualitatively the observed sublinear dependence of theTHz amplitude on the excitation intensity.

In contrast, in the high-excitation-density condition, nP > 1011 cm−2,which corresponds to the 0.2 mm spot case and the excitation > 6 mW inthe 0.8 mm spot case in Fig. 30, n0 + nP is dominated by nP and increasesrapidly, and Eeff drops steeply, resulting in a strong saturation of THz waveamplitude. In such a case, however, the recovery current during each pulsecan not be neglected, since ∆V becomes large (the potential barrier for thethermionic emission becomes small), and the present simple calculation doesnot predict correctly the experimental results. Possibly, as observed experi-mentally, due to the instantaneous recovery, the THz amplitude may increasesgradually with the increase of the excitation intensity.

6 Summary

In this Chapter, recent topics on the time-domain THz-emission studies ofbulk and quantum semiconductor structures were presented, including thediscussion on the applications of such structures as the emitter of THz waves.

In Sect. 2, the nonequilibrium carrier transport in DC-biased bulk GaAswere discussed quantitatively on the basis of the time-domain THz-emissionexperiment compared with the ensemble Monte Carlo calculations. As a re-sult, it was proven that the finiteness of sample thickness has to be takeninto account in the analysis of the THz data. Furthermore, from the dis-cussion on the steady-state carrier velocity, we showed that the contributionof hole transport becomes significant for a DC-bias higher than 20 kV/cm.It was then concluded that, when actual experimental conditions (samplegeometry, photoexcitation condition, etc.) are properly taken into account,Monte Carlo calculations give a very good description of the transient carriertransport determined by the THz spectroscopy.

Page 38: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

154 Yutaka Kadoya and Kazuhiko Hirakawa

Section 3 was devoted to the THz wave emissions from semiconductor su-perlattices. In particular, we focused on a long-standing question, i.e., Blochgain in semiconductor superlattices. To answer this question experimentally,we have showed that the time-domain THz spectroscopy is very useful indetermining gain/loss spectra in the THz regime. By noting that the time-domain THz emission spectroscopy inherently measures the step responseof the electron system to the bias electric field, the obtained THz spectrawere compared with the high-frequency conductivities predicted for minibandtransport. Excellent agreement between theory and experiment strongly sup-ports that the THz gain due to Bloch-oscillating electrons persists at least upto 1.7 THz. It was also shown that Zener tunneling into the second minibandsets the high-frequency limit to the THz gain for the samples studied here.

Radiation of THz waves from high-Q semiconductor microcavities, wherethe quantum-well excitons and the cavity-mode photons are strongly cou-pled, was presented in Sect. 4. A systematic investigation on the frequencyand amplitude of the THz waves showed clearly that the observed THz wavesare radiated from the coherent oscillation of the exciton population, and thatthe radiation can be viewed as a kind of second-order nonlinear response. Inaddition, the transition from strong- to weak-coupling regime with the de-crease of the cavity Q-value was demonstrated as the decay of the oscillationin the THz wave.

In Section 5, we discussed the radiation of THz waves from low-Q bulksemiconductor microcavities containing a bulk GaAs layer. It was shown that,in the low-excitation-density regime, < 1010 cm−2 photons per pulse, the useof an optical microcavity enhances the THz intensity in comparison withthe bare semiconductor surfaces, owing to the efficient use of the excitationpulses as well as to the reduction of the reflection of the excitation pulses atthe surface. In addition, some results of a numerical estimation were shownto point out that the screening of the surface DC field by the accumulatedphotocarriers can cause sublinear dependence of the THz amplitude on theexcitation intensity, even in the low-excitation-density regime.

As demonstrated in this Chapter, the time-domain THz-emission studiesprovide rich information on the temporal as well as spatial dynamics of pho-toexcited carriers in semiconductors on the subpicosecond time scale, wherethe nonstationary and/or coherent effects play a significant role. In additionto the excellent time resolution, the technique has a feature that the obtainedwaveform is related directly to the photocarrier dynamics. Hence, the tech-nique is unique and useful, and expected to be used in the future even morewidely. On the other hand, the use of various semiconductor structures isexpected to lead to the improvement or the development of the devices forthe emission, detection, and control of THz waves, which may be required inthe future high-performance, low-cost, compact, and functional THz systems.Further investigations in such a direction are strongly desired.

Page 39: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 155

References

[1] W. Sha, J. Rhee, T. Norris, W. Schaff: IEEE J. Quantum Electron. 28, 2445(1992) 117

[2] A. Iverson, G. Wysin, D. Smith, A. Redondo: Appl. Phys. Lett. 52, 2148 (1988)117

[3] B. Hu, E. de Souza, W. Knox, J. Cunningham, M. Nuss, A. Kuznetsov,S. Chuang: Phys. Rev. Lett. 74, 1689 (1995) 117

[4] H. Roskos, M. Nuss, J. Shah, K. Leo, D. Miller, A. Fox, S. Schmitt-Rink,K. Kohler: Phys. Rev. Lett. 68, 2216 (1992) 118

[5] P. Planken, M. Nuss, I. Brener, K. Goossen, M. Luo, S. Chuang, L. Pfeiffer:Phys. Rev. Lett. 69, 3800 (1992) 118

[6] C. Waschke, H. Roskos, R. Schwedler, K. Leo, H. Kurz, K. Kohler: Phys. Rev.Lett. 70, 3319 (1993) 118, 127, 132

[7] P. Bolivar, F. Wolter, A. Muller, H. Roskos, H. Kurz, K. Kohler: Phys. Rev.Lett. 78, 2232 (1997) 118

[8] Y. Yamashita, A. Endoh, K. Shinohara, K. Hikosaka, T. Matsui, S. Hiyamizu,T. Mimura: IEEE Electron Dev. Lett. 23, 573 (2002) 119

[9] A. Leitenstorfer, S. Hunsche, J. Shah, M. Nuss, W. Knox: Phys. Rev. Lett.82, 5140 (1999) 119, 147

[10] A. Leitenstorfer, S. Hunsche, J. Shah, M. Nuss, W. Knox: Phys. Rev. B 61,16 642 (2000) 119

[11] M. Abe, S. Madhavi, Y. Shimada, Y. Otsuka, K. Hirakawa: Appl. Phys. Lett.81, 679 (2002) 120, 121, 122, 123, 124, 125

[12] Q. Wu, X.-C. Zhang: IEEE J. Sel. Top. Quantum Electron. 2, 693 (1996) 121[13] Q. Wu, X.-C. Zhang: Appl. Phys. Lett. 70, 1784 (1997) 121[14] A. Leitenstorfer, S. Hunsche, J. Shah, M. Nuss, W. Knox: Appl. Phys. Lett.

74, 1516 (1999) 121[15] G. Gallot, J. Zhang, R. McGowan, T.-I. Jeon, D. Grischkowsky: Appl. Phys.

Lett. 74, 3450 (1999) 121[16] E. Palik: Handbook of Optical Constants of Solids 2 (Academic, New York

1991) 121[17] K. Tomizawa: Numerical Simulation of Submicron Semiconductor Devices

(Artech House, Boston 1993) 122[18] C. Jacoboni, P. Lugli: The Monte Carlo Method for Semiconductor Device

Simulation (Springer, Berlin, Heidelberg 1989) 122[19] F. Bloch: Z. Phys. 52, 555 (1928) 125[20] C. Zener: Proc. R. Soc. London Ser. A 145, 523 (1934) 125[21] L. Esaki, R. Tsu: IBM J. Res. Dev. 14, 61 (1970) 126[22] J. Feldmann, K. Leo, J. Shah, D. Miller, J. Cunningham, T. Meier, G. von

Plessen, A. Schulze, P. Thomas, S. Schmitt-Rink: Phys. Rev. B 46, 7252 (1992)127, 132

[23] T. Dekorsy, P. Leisching, K. Kohler, H. Kurz: Phys. Rev. B. 50, 8106 (1994)127, 132

[24] F. Loser, Y. Kosevich, K. Kohler, K. Leo: Phys. Rev. B 61, 13373 (2000) 127,132

[25] Y. Shimada, K. Hirakawa, S.-W. Lee: Appl. Phys. Lett. 81, 1642 (2002) 127,129, 130, 131, 132, 137

[26] G. Bastard, R. Ferreira: Cr. Acad. Sci. II 312, 971 (1991) 127, 132

Page 40: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

156 Yutaka Kadoya and Kazuhiko Hirakawa

[27] Y. Shimada, K. Hirakawa, M. Odnoblioudov, K. Chao: Phys. Rev. Lett. 90,46806 (2003) 127, 128, 131, 133, 134, 135, 136, 137

[28] R. Kersting, K. Unterrainer, G. Strasser, H. Kauffmann, E. Gornik: Phys. Rev.Lett. 79, 3038 (1997) 129

[29] H. Kroemer: On the nature of the negative-conductivity resonance in a super-lattice bloch oscillator, arXiv: cond-mat/0007428 (2000) 132

[30] H. Willenberg, G. H. Dohler, J. Faist: Phys. Rev. B 67, 085315 (2003) 132[31] S. Allen, et. al.: Semicond. Sci. Technol. 7, B1 (1992) 133[32] M. Helm: Semicond. Sci. Technol. 10, 557 (1995) 133[33] S. Ktitorov, G. Simin, V. Sindalovskii: Sov. Phys. Solid State 13, 1872 (1971)

134[34] B. Rosam, K. Leo, M. Gluck, F. Keck, H. Korsch, F. Zimmer, K. Kohler: Phys.

Rev. B 68, 125301 (2003) 137, 138[35] Y. Shimada, N. Sekine, K. Hirakawa: Appl. Phys. Lett. (2004) in press 137,

138, 139[36] A. Sibille, J. Palmier, F. Laruelle: Phys. Rev. Lett. 80, 4506 (1998) 138[37] M. Helm, W. Hilber, G. Strasser, R. D. Meester, F. Peeters, A. Wacker: Phys.

Rev. Lett. 82, 3120 (1999) 138[38] Y. Kadoya: Optical Properties of Low-Dimensional Materials 2 (World Scien-

tific, Singapore 1998) Chap. 7 140[39] C. Weisbuch, M. Nishioka, A. Ishikawa, Y. Arakawa: Phys. Rev. Lett. 69, 3314

(1992) 140[40] T. Norris, J. Rhee, C. Sung, Y. Arakawa, M. Nishioka, C. Weisbuch: Phys.

Rev. B 50, 14663 (1995) 141[41] H. Cao, J. Jacobson, G. Bjork, S. Pau, Y. Yamamoto: Appl. Phys. Lett. 66,

1107 (1995) 141[42] J. Jacobson, S. Pau, H. Cao, G. Bjork, Y. Yamamoto: Phys. Rev. A 51, 2542

(1995) 141[43] H. Wang, J. Shah, T. Damen, W. Jan, J. Cunningham, M. Hong, J. Mannaerts:

Phys. Rev. B 51, 14713 (1995) 141[44] G. Bongiovanni, A. Mura, F. Quochi, S. Gurtler, J. Staehli, F. Tassone,

R. Stanley, U. Oesterle, R. Houdre: Phys. Rev. B 97, 7084 (1997) 141[45] M. Koch, J. Shah, T. Meier: Phys. Rev. B 57, R2049 (1998) 141[46] H. Wang, H. Hou, B. Hammons: Phys. Rev. Lett. 81, 3255 (1998) 141[47] Y. Kadoya, K. Kameda, M. Yamanishi, T. Nishikawa, T. Kannari, T. Ishihara,

I. Ogura: Appl. Phys. Lett. 68, 281 (1996) 141, 145[48] Y. Hokomoto, Y. Kadoya, M. Yamanishi: Appl. Phys. Lett. 74, 3839 (1999)

142[49] M. Kusuda, S. Tokai, Y. Hokomoto, Y. Kadoya, M. Yamanishi: Physica B 272,

467 (1999) 142[50] V. Savona, F. Tassone: Solid State Commun. 95, 673 (1995) 145[51] P. Planken, M. Nuss, W. Knox, D. Miller, K. Goossen: Appl. Phys. Lett. 61,

2009 (1992) 146[52] X.-C. Zhang, B. Hu, J. Darrow, D. Auston: Appl. Phys. Lett. 56, 1011 (1990)

147[53] S. Chuang, S. Schmitt-Rink, B. Greene, P. Aseta, A. Levi: Phys. Rev. Lett.

68, 102 (1992) 147

Page 41: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Terahertz Radiation from Bulk and Quantum Semiconductor Structures 157

[54] T. Sakurada, Y. Kadoya, M. Yamanishi: Jpn. J. Appl. Phys. 41, L256 (2002)147, 150, 151, 152

[55] H. Casey, M. Panish: Heterostructure Lasers (Academic, New York 1978) p. Pt.A 153 150

[56] S. Sze: Physics of Semiconductor Devices, 2nd ed. (Wiley, New York 1981)Chap. 5 152

Page 42: [Topics in Applied Physics] Terahertz Optoelectronics Volume 97 || Terahertz Radiation from Bulk and Quantum Semiconductor Structures

Index

acceleration, 117, 119, 122anticrossing, 140autocorrelation-type interferometer,

129

Bloch oscillation, 125Bloch oscillator, 126, 132Boltzmann transport, 134built-in potential, 147

carrier accumulation, 151cavity-mode, 140cavity-polariton, 139conductivity, 134

deceleration, 117, 119dephasing, 144dispersion curve, 140distributed Bragg reflector (DBR), 142

exciton, 128, 140

gallium arsenide (GaAs), 120

hole transport, 124

inversion symmetry, 141

microcavity, 139miniband, 126molecular beam epitaxy, 127, 148

momentum relaxation time, 134Monte Carlo calculation, 119

nonstationary carrier transport, 119normal mode oscillation (NMO), 140

optical rectification, 141oscillator strength, 140overshoot, 119, 122

polariton, 139

quantum well (QW), 118, 139quantum-confined Stark effect (QCSE),

142

Rabi frequency, 140relaxation time, 134

scattering, 122, 123, 132screening, 151steady-state velocity, 124strong coupling, 140superlattice (SL), 118

THz gain, 125transient carrier transport, 119

Wannier–Stark (WS) ladder, 126

Zener tunneling, 136