Sunteți pe pagina 1din 12

IEICE TRANS. ELECTRON., VOL.E83C, NO.

6 JUNE 2000

938

INVITED PAPER

Special Issue on Advanced Optical Devices for Next Generation Photonic Networks

InP/InGaAs Uni-Traveling-Carrier Photodiodes


Tadao ISHIBASHIa) , Tomofumi FURUTA , Hiroshi FUSHIMI , Satoshi KODAMA ,
Hiroshi ITO , Tadao NAGATSUMA , Naofumi SHIMIZU ,
and Yutaka MIYAMOTO , Regular Members

SUMMARY This paper reviews the operation, design, and


performance of the uni-traveling-carrier-photodiode (UTC-PD).
The UTC-PD is a new type of photodiode that uses only electrons as its active carriers and its prime feature is high current
operation. A small signal analysis predicts that a UTC-PD can
respond to an optical signal as fast as or faster than a pin-PD.
A comparison of measured pulse photoresponse data reveals how
the saturation mechanisms of the UTC-PD and pin-PD dier.
Applications of InP/InGaAs UTC-PDs as optoelectronic drivers
are also presented.
key words: photodiode, photoreceiver, InP, InGaAs

1.

Introduction

The advent of optical ampliers has changed the device


requirement of optical signal detection in high bit-rate
ber-optic communication and optoelectronic measurement systems. In assessing the performance of a photodiode (PD), bandwidth and quantum eciency have
been the basic gures-of-merit. In addition to these attributes, high output (large signal operation) has now
been very important. This is because, as the signal
clock rate or microwave carrier frequency increases, a
photoreceiver consisting of an optical pre-amplier and
a high power photodiode (Fig. 1) can produce better
characteristics than a conventional photoreceiver with
electrical post-ampliers. In a high bit-rate time division multiplexing (TDM) system, for example, this new
photoreceiver conguration without electronic ampliers is benecial in terms of system simplicity and input
sensitivity [1]. Similarly, eliminating high-gain ampliers can reduce complexity of optoelectronic ICs with
optical inputs. Microwave photonics systems generally
need the optical generation of micro/millimeter-wave
signals, so wide linearity and high saturation output are
the basic requirements of PDs. In the millimeter-wave
range, an optoelectronic transmitter without ampliers
can simplify system conguration and realize a wider
range of applications.
Manuscript received January 31, 2000.
The authors are with NTT Photonics Laboratories,
Atsugi-shi, 243-0198 Japan.

The author is with NTT Telecommunications Energy


Laboratories, Atsugi-shi, 243-0198 Japan.

The authors are with NTT Network Innovation Laboratories, Yokosuka-shi, 239-0847 Japan.
a) E-mail: ishi@aecl.ntt.co.jp

The large signal/high power operation of PDs is


thus a key goal. Up to present, saturation characteristics of conventional pin-PDs with microwavemodulated optical signals have been investigated [2], [3].
1-dB-compression output was reported to be 12 dBm
at 20 GHz for a device biased at 4 V [3]. Due to the
tradeo between bandwidth (f3dB ) and available current, however, the maximum output determined by the
junction area is roughly proportional to (1/f3dB )2 . This
kind of trend, generally followed by pin-PDs, is a constraint in the millimeter-wave frequency range. In order to enhance output, the velocity-matched (VM) distributed photodetector with multi-stage PDs (or metalsemiconductor-metal detectors) has been also studied
[4], [5]. VM photodetectors can be free from the limitation of the CR time constant and their greater lightabsorption volume allows a higher saturation current,
although the saturation itself still remains. In pulse
photoresponse measurements, a peak current of 56 mA
was obtained from a VM-MSM photodetector (f3dB of
49 GHz) made of AlGaAs/GaAs [5].
The saturation in photocurrent originates from the
eld modulation by space charges in the diode deple-

Fig. 1 A photoreceiver consisting of an optical pre-amplier


and a high power photodiode (a) and a conventional photoreceiver with electrical post-ampliers (b).

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

939

Fig. 2

Schematic band diagrams for a UTC-PD and a pin-PD.

tion layer, and is usually frequency dependent. Because holes have a much lower drift velocity than electrons, their transport dominates the space charge eect
and saturation behavior in a pin-PD. The uni-travelingcarrier photodiode (UTC-PD) was designed to oer
higher operation current densities [6]. The benets of
UTC-PDs are created from their use of electrons as the
only active carriers. Here, hole transport does not directly aect the diode response and output saturation
mechanisms. InP/InGaAs UTC-PDs have been fabricated with various diode dimensions and demonstrated
to produce signicantly higher output.
This paper reviews the operation, design, and performance of the UTC-PD. We discuss how its photoresponse diers from that of the pin-PD. The application
of InP/InGaAs UTC-PDs as optoelectronic drivers is
also presented.
2.

Uni-Traveling-Carrier Photodiode and pin


Photodiode

Band diagrams are schematically shown for a UTCPD and a pin-PD in Fig. 2. The active layer of the
UTC-PD consists of a neutral (p-type) light absorption layer and a widegap (depleted) carrier-collection
layer. Carriers are photogenerated in the absorption
layer, and minority electrons diuse/drift into the collection layer. Here, the diusion block layer gives
electrons unidirectional motion toward the collection
layer. The quasi-eld formed by the band gap grading and/or doping grading is very eective in reduc-

ing electron traveling time. Since holes are the majority carriers, their transport behaves depending on the
electron current. Consequently, the photoresponse of a
UTC-PD is determined by the electron transport in the
whole structure. In a typical UTC-PD structure with
similar absorption layer thickness WA and collection
layer thickness WC , the absorption layer traveling time
A is dominant because the diusive velocity is usually smaller than drift velocity in the collection layer
with high electric eld. As in the case of heterostructure bipolar transistors [7], electrons represent velocity overshoot in the collection layer, where the velocity
(vos 4 107 cm/s) is far greater than hole saturation
velocity (vs 5 106 cm/s).
In a pin-PD, on the other hand, both electrons
and holes in the depleted absorption layer contribute
to the response. When electrons and holes travel independently, the output response is sum of both current
components. Here, due to the signicant dierence in
carrier velocities, the 3-dB-down bandwidth is determined by the hole transport.
Since the carrier generation and collection are separated in space, the total carrier travel distance of a
UTC-PD is relatively long. Lets compare a pin-PD
with the absorption layer thick ness of WA to a UTCPD with WA = WC . Under uniform carrier generation
over the absorption layer, the average carrier traveling
distance is 0.5 WA in the pin-PD, while it is 1.5 WA
in the UTC-PD. In spite of this drawback, the total
carrier travel time of a UTC-PD, which is determined
by the electron transport, can be similar to or a little
shorter than that of a pin-PD. This condition is realized in InP/InGaAs UTC-PDs designed for a long wavelength operation. Here, the minority-electron mobility
e (diusivity De ) is a key parameter. We should note
that the eective electron diusion velocity, 3De /WA
in InGaAs, can be higher than the hole saturation velocity vs . At a limiting case where the collection layer
travel time C (= WC /vos ) dominates (or WA = 0), the
bandwidth f3dB is simplied to 2.8/(2Celectron ) for a
UTC-PD, while it is given by 3.5/(2Chole ) for a pinPD (Chole = WC /vs ). Due to the large dierence in
carrier velocities, the UTC design can nally provide a
shorter C and superior frequency response.
3.

Output Saturation Induced by the Space


Charge Eect

With increasing photocurrent, mobile charge density


increases in the depletion layer of a photodiode and
modulates the eld prole, which then presents nonlinear response and output saturation. The UTC-PD and
pin-PD exhibit dierent space charge eects [6]. This
reectes on the critical current densities Jmax at which
the eld modulation starts to degrade the diode operations. We consider a steady state assuming, for simplicity, constant carrier saturation velocities, vselectron

IEICE TRANS. ELECTRON., VOL.E83C, NO.6 JUNE 2000

940

Fig. 4

Fig. 3 Charge distribution, eld, and band bending at high


carrier injection levels.

and vshole , and roughly estimate Jmax . The changes in


band bending at high carrier injection levels are shown
for both structures in Fig. 3. The UTC-PD shows a
uniform negative charge distribution, while the pin-PD
shows a triangular positive charge distribution (ignoring electron charge). The eld intensity is lowered on
the p-type layer side in the UTC-PD, and on the n-type
layer side in the pin-PD. Finally, the elds can reach
electron and E hole below that
the critical electric elds EC
C
the saturation velocities are not maintained. The corresponding carrier densities nC and pC that yield such
electric elds are simply calculated by solving Poissons
equation as follows
electron W ]/(qW 2 ) (1a)
nC = 2[Vbias + Vbi EC
C
C
hole W ]/(qW 2 ),
pC = 6[Vbias + Vbi EC
C
C

(1b)

where is dielectric constant, Vbias is applied voltage,


Vbi is built-in voltage, and WC is the depletion width.
When the terms that include EC are ignored, the resulting ratio between Jmax values can be approximated
as
Jmax (U T C)/Jmax (pin)
= qnC vselectron /qpC vshole (vselectron /vshole )/3.
(2)
In a pin-PD, the actual hole current component is
qpC vshole /2. Assuming the hole saturation velocity
of 5 106 cm/s in InGaAs and the electron overshoot velocity of 4 107 cm/s in InP, Eq. (2) gives

Layout of the UTC-PD model for small-signal analysis.

Jmax (U T C)/Jmax (pin) 3. In spite of the lower


critical carrier density, a UTC-PD can be driven at
three times higher current. Actually, at typical bias
hole W is not negligible. For
voltages, the term EC
C
hole
example, for EC
= 50 kV/cm and WC = 0.4 m,
hole W is 2 V, while E electron = 5 kV/cm results
EC
C
C
electron W
in EC
C = 0.2 V. At a small negative bias,
Vbias Vbi + 2 = 1.25 V, therefore, such a pin-PD
does not work properly at velocity saturation for holes.
On the other hand, a UTC-PD can operate well even at
a positive bias voltage. The dierence will be further
discussed in Sect. 6 in connection with experimental results.
4.

UTC-PD Model and Small-Signal Photoresponse

4.1 Modeling of UTC-PDs


One theoretical analysis of the small-signal frequency
photoresponse of UTC-PDs used a drift-diusion model
[8]. The layout is shown in Fig. 4. In a small signal
case, transport of minority electron in the absorption
layer is treated as being independent of hole transport (similarly to that for a bipolar transistor). By
solving the continuity equation for minority electron
density n(x, ), including the carrier generation term
Gexp(jt), electron current Je (x) is obtained as the
sum of diusion and drift components.
Since majority hole current Jh (x) is related to electron current through the current continuity, total particle current Je (x) +Jh (x) can be expressed as a function
of Je (x)

jR Je (x)
[Je (x) + Jh (x)] =
x
1 + jR x

(3)

where R is dielectric relaxation time / in the absorption layer. In deriving Eq. (3), we ignore the hole
diusion current and take Jh (x) to be E(x). Equation (3) just indicates that the continuity of particle
current can be maintained at a frequency  1/R
and, to achieve this condition, the absorption layer doping has to be high enough. On the other hand, electron
current in the collection layer is regarded as current
induced in a parallel capacitor. The frequency photoresponse is the total (short circuit) photocurrent Jtot

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

941

Fig. 5 Electron density distribution in the absorption layer of


a UTC-PD with absorption layer thickness of 0.2 m. Minority
electron mobility of 4000 cm2 /Vs is assumed.

through the diode


 WA
 WA +WC
1
1
[Je (x)+Jh (x)]dx+
Je (x)dx
Jtot =
W 0
W WA




1 WA
jR
Je (x)
=
Je (WA ) 1
1
dx
W 0
1+jR
Je (WA )


sin(C /2)
WC
Je (WA )
exp(jC /2)
+
W
C /2
(4)
where W is the sum of WA and WC , C is the carrier traveling time in the collection layer (WC /vos ), and
Je (WA ) is electron current at the interface of absorption/collection layers (x = WA ). Finally, by calculating
the electron distribution n(x, ) and Je (WA ), we can
evaluate the response Jtot from Eq. (4).
4.2 Calculated Photoresponse
First, we examine the electron density distribution
n(x, = 0) in the absorption layer at steady state conditions. In deriving the distribution, the boundary condition at the absorption/collection layer interface was
determined in such a way that the diusion current
De (n/x) is given by the thermionic emission current
qvth n(WA ), where vth is thermionic emission velocity
(2.5107 cm/s in InGaAs). To study the eect of quasield Fqf , potential variation over the absorption layer
( = Fqf WA ) is changed as a parameter (see Fig. 5). At
= 0, where only diusion transport exists, n(x) has a
parabolic shape. As increases, electron density is signicantly reduced and the distribution shifts to the collection layer. The lower electron density means higher
electron velocity and shorter travel time. We also note
that the electron density is not very low at the absorption/collection layer interface, indicating that the
velocity is limited by the thermionic emission process.
This is characteristic of high mobility materials and

Fig. 6 Variation of f3dB with absorption layer thickness WA


calculated for UTC-PDs (WA = WC ) and pin-PDs.

suggests the importance of the boundary condition in


accurately modeling UTC-PDs.
Figure 6 shows the variation of f3dB with absorption layer thickness WA calculated for InP/InGaAsP
UTC-PDs (WA is taken to be equal to WC ). Here
the eect of R is negligible for the doping level of
p = 1 1018 /cm2 assumed. It is found that f3dB of
the UTC-PDs changes roughly as 1/WA2 , which is in
contrast to the variation in a pin-PD, f3dB 1/WA
(shown by dashed line). The gure also shows that
the incorporation of a quasi-eld signicantly increases
f3dB . Even for the small of 50 mV (E = 2.5 kV/cm for
WA = 0.2 m), f3dB can be doubled. This is because
the drift velocity e (/WA ) is higher than diusion
velocity 3De /WA . For WA = 0.2 m, f3dB increases
from 68 GHz ( = 0) to 160 GHz ( = 50 mV).
In the above calculation, the response was calculated by taking account of both absorption and collection layer travel times A and C . However, A is
usually much longer than C as far as WA is similar to
WC . In such a case, the response is well approximated
by Je (WA ). Furthermore, in a pure diusion case (no
quasi-eld), the response is simplied to


JDC
W2
WA
;
A = A +
Je =
(5)
1 + jA
3De
vth
where JDC is average photocurrent qGWA . The rst
term in A is the carrier travel time determined by the
diusive motion of minority electrons. As WA falls, the
second term in (5) becomes dominant [9]. Variation of
A as a function of WA is plotted in Fig. 7 with absorption layer doping level as a parameter. Electron mobilities are taken to be e = 7000, 5000 and 3200 cm2 /Vs,
respectively, for InGaAs absorption layer doping levels
of p = 2 1017 , 1 1018 and 5 1018 /cm3 . In calculating A , the eect of dielectric relaxation was not
considered for simplicity, and instead, curves are drawn
for A in the region A > 2R . Clearly, a lower acceptor doping level is preferable to minimize (maximize)
A (f3dB ), since diusivity is proportional to mobility,

IEICE TRANS. ELECTRON., VOL.E83C, NO.6 JUNE 2000

942

Fig. 7
of WA .

Calculated variation of A for UTC-PDs as a function

Table 1 Epitaxial layer structure of an InP/InGaAs UTC-PD


with compositional grading combined with planar doping [11].
Layer
Thickness (nm) Doping (cm3 ) EG (eV)
p+ -InGaAs Cap
50
3.0E19
20
2.0E19
0.85
p+ -InGaAsP Di.Barrier
p-InGaAs Absorber
220
1.0E18
i-InGaAs Spacer
8

i-InGaAsP Spacer
16

1.0
i-InP Spacer
6

7
1.0E18
n+ -InP Cli.
n-InP Collector
263
1.0E16
50
5.0E18
n+ -InP Subcollec.-2
n+ -InGaAs Etch Stop
10
1.5E19
500
1.5E19
n+ -InP Subcollec.-1
i-InGaAs Etch stop
10

(S. I. InP Sub)

De = kT e /q. The inuence of the maximum diusive velocity is not negligible as WA decreases. For
WA values larger than 100 nm, A tends to change as
WA2 indicating classical diusive transport. However,
this diusive mode becomes less valid as WA falls; A
changes more gradually as WA /vth determines it, not
the electron diusivity (doping level).
5.

Fabrication and Characterization of


InP/InGaAs UTC-PDs

InP/InGaAs UTC-PDs have been fabricated with MOVPE (metal-organic vapor-phase-epitaxy) grown epitaxial materials. A typical structure uses p+ -InGaAs
and n -InP as the absorption- and collection-layers,
respectively (Table 1). In layer design, an important
point is the interface between absorption/collection
layers. Due to the conduction band discontinuity,
240 meV at the InP/InGaAs heterointerface, the InP
potential spike can produce current blocking unless
the band prole is designed carefully. This problem can be most eectively eased by adopting compositional grading combined with planar doping [10],

Fig. 8 Simulated band proles for UTC-PD structures with


and without compositional grading.

[11]. Additional intermediate layers of i-InGaAs/iInGaAsP(EG = 1.0 eV)/i-InP/n+ -InP are inserted between the p+ -InGaAs absorption- and n-InP collectionlayers. The thickness of each layer and the doping level
for n+ -InP are optimized by numerical simulation so
that a smooth conduction band change is realized at
the junction voltage used. Figure 8 presents simulated
band proles for structures with and without compositional grading. It is clearly seen that the use of the
i-InGaAsP layer can reduce not only the potential spike
height, but also the potential notch depth.
A double-mass structure is used for fabricating
InP/InGaAs UTC-PDs. In processing back-illuminated
planer devices, conventional metal lift-o for Ohmic
contacts, wet chemical etching, and electrode-isolation
by polyimide lm have been used. Edge-illuminated
waveguide devices have also been reported [12]. The
responsivity is generally the same as that of a pin-PD
as far as each absorption layer thickness is equal. When
applying UTC-PDs to photoreceivers, they are integrated with a 50- coplanar waveguide (CPW) line,
a bias capacitor, and a matched resistor. Figure 9 illustrates a planar device used for pulse-photoresponse

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

943

Fig. 9 Device geometry of a planar device used for pulsephotoresponse characterization by means of the electro-optic
sampling (EOS) technique.

characterization by means of the electro-optic sampling


technique. The test diode is located at the center of the
chip and the signal output is connected to two 50-
CPWs patterned on the semi-insulating InP substrate.
To perform pump-probe measurements at a wavelength
of 1.55 m, a CdTe probe chip is placed on a CPW line.
The eective load impedance of a diode is 25 in the
EOS measurement. If the device junction diameter is
46 m, the CR time constant can be reduced to around
500 fs, making it possible to observe the response resulting from the carrier dynamics in the device.
6.

Photoresponse of UTC-PDs

6.1 Dierence between a UTC-PD and a pin-PD


Pulse-photoresponse waveforms for a UTC-PD with
various optical inputs (0.2 to 2.0 pJ/pulse, F W HM =
390 fs) are shown in Fig. 10(a) [10]. The absorption and
collection layers are 220 and 300 nm-thick, respectively,
and the diode area is 20 m2 . Bias voltage is 2 V.
When the input power increases, the output intensity
increases almost linearly for inputs up to 1.0 pJ/pulse.
The peak voltage of 0.75 V corresponds to a current
value of 30 mA. With higher inputs, the output starts
to saturate and pulse widening occurs. Here, the fall
time changes little even when output saturation occurs.
This indicates that the saturation takes place at currents exceeding a certain level and that the operation
can return to the linear response mode. For comparison, output waveforms for a pin-PD (fabricated in the
same way as the UTC-PD) with absorption layer thickness WA = 300 nm is shown in Fig. 10(b) [10]. The output waveforms dier greatly from those of the UTC-PD,
and two current components are observed. The initial sharp peak comes primarily from the electron cur-

Fig. 10 (a) Pulse-photoresponse waveforms for a UTC-PD


with WA = 220 nm and WC = 300 nm, and (b) for a pin-PD
with WA = 300 nm.

rent in the depletion layer. As the input increases, tail


currents extend signicantly with longer decay times,
indicating nonlinearity and saturation. Although two
current components are not resolved numerically, saturation commences at outputs (for a tail component) of
the order of 0.10.15 V as seen in the waveform for the
0.4 pJ/pulse.
Output saturation is basically induced by carrier
space charges in both photodiode structures. The saturation point can be described in terms of the critical
charge and current density for a given voltage (Sect. 3).
Calculated critical current densities, Jmax (U T C) and
Jmax (pin) are plotted as functions of operation voltage
in Fig. 11. In the pulse measurements, the operation
voltage moves along a load line (dashed line, drawn
for a 20 m2 device with a 25- load), so the intersection gives the saturation point. The critical outputs
are 0.88 V and 0.20 V for the UTC-PD and pin-PD, respectively. These values show approximate coincidence
with the ones experimentally observed. After the onset of saturation, as we have seen, the UTC-PD and
pin-PD behave dissimilarly, implying that their carrier
dynamics are qualitatively dierent. In the UTC-PD,
current blocking suppresses electron injection into the
collection layer and is the cause of current saturation.
On the other hand, the bi-traveling-carrier character-

IEICE TRANS. ELECTRON., VOL.E83C, NO.6 JUNE 2000

944

Fig. 11 Calculated critical current densities, Jmax (U T C) and


Jmax (pin) as functions of operation voltage.

Fig. 13 3-dB-down frequency (f3dB ) evaluated from the


Fourier transform of the pulse waveform in Fig. 10.

eling time. The measured f3dB s at low outputs (solid


circles) agree fairly well with the calculated results. At
high output conditions, however, the evaluated f3dB s
(open circles) are much higher than those at low outputs. This will be explained below.
6.3 Self-Induced Field in the Absorption Layer of a
UTC-PD

Fig. 12 Measured and reported f3dB values against absorption


layer width WA . The dashed line (f3dB vs. WA ) is calculated
for UTC-PDs by using Eq. (4).

istic of the pin-PD makes the carrier dynamics very


complicated [13]. The presence of the long tail implies
that the eld screening in the depletion layer initiated
by hole storage is suppressing the carrier motions.
6.2 Variation of Bandwidth with Absorption Layer
Width
Measured and reported f3dB values are plotted against
absorption layer width WA in Fig. 12 for UTC and pinPDs [9], [14], [17]. f3dB for pin-PDs changes as 1/WA ,
while for UTC-PDs, approximately as 1/WA2 (some
scatter in the data is due to dierent absorption layer
doping levels). Thicknesses of the collection layers
WC s range from 200 nm to 230 nm in UTC-PDs, where
f3dB s are little inuenced by the change in WC . It is
found that the UTC design is advantageous in achieving
f3dB values higher than around 100 GHz. The dashed
line is the variation in f3dB calculated by using Eq. (4).
Here, deviation of the curve from 1/WA2 is primarily
due to the nite value of the thermionic emission velocity, not the contribution of the collection layer trav-

When carefully looking at the output waveforms in


Fig. 10(a) (UTC-PD), the output pulse sharpening
is observed. As the input increases from 0.2 to
1.0 pJ/pulse, the FWHM falls from 3.5 to 2.8 ps. This
variation can be more clearly seen in the 3-dB-down
frequency (f3dB ) evaluated from the Fourier transform
of the pulse waveform (Fig. 13). f3dB of the UTC-PD
increases by about 50% with inputs from 63 GHz to
94 GHz, and decreases thereafter due to output saturation. In contrast, f3dB monotonically decreases in the
pin-PD [10]. The self-induced eld Eind in the absorption layer [15] has explained the signicant increase in
the f3dB of the UTC-PD. The origin of Eind is the drift
current Jhole (x) of majority holes
Jhole (x) p Eind (x)

(6)

that must ow to maintain the current continuity,


where p is the conductance of the absorption layer
(= qp0 h ). Jhole (x) is zero at the absorption/collection
layer interface, and linearly increases (under uniform
carrier generation) toward the p-contact layer side up to
the diode operation current Jop . Eind does not depend
on the absorption layer thickness WA , while the diusive velocity drops with increasing WA , thus this eect
is more intense for wider WA values. Higher absorption
layer doping can suppress this eect. For a moderate
doping level, p0 = 1018 /cm3 in the present UTC-PD,
Eind can be high enough to accelerate electrons. For
example, at Jop = 100 kA/cm2 , the maximum Eind is
calculated to be 5.6 kV/cm. With minority electron
mobility of 5000 cm2 /Vs, this Eind results in a drift

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

945

Fig. 14 Variation of pulse peak current Ipeak with optical input


for a UTC-PD with WA = 220 nm and WC = 300 nm.

Fig. 15 Pulse peak current Ipeak vs. operation voltage (Vop =


Vbias + Vout ) characteristics for a UTC-PD with WA = 220 nm
and WC = 300 nm.

velocity vd of 2.8 107 cm/s which is higher than the


pure diusive velocity 3De /WA = 1.8 107 cm/s, and
comparable to the thermionic emission velocity. This
is why a signicant increase in f3dB was observed. Although the self-induced eld is a kind of nonlinear eect
peculiar to UTC-PDs, it presents no disadvantages provided f3dB is designed to be higher enough than signal
frequency.
6.4 High Output Operation
Saturation output of a UTC-PD is a function of many
parameters including bias voltage and device size. Figure 14 shows how the peak current (Ipeak ) vs. input
curves change with bias Vbias (1 to 4 V) for a UTCPD with a diode area of 40 m2 [9]. With increasing
negative bias Vbias , the saturation level increases, but
the saturation tendency is gradual at large Vbias values
(3 and 4 V). This is attributed to the fact that electron velocity decreases with bias (as described in the
next subsection) that leads to an increase in the level
of the space charge density. Nevertheless, the linearresponse region extends to about 80 mA (200 kA/cm2 )
at Vbias of 3 V. Here, the FWHM of the output pulse
was 3.3 ps, which is similar to that for the 20 m2 device
shown in Fig. 10(a).
The saturation is strongly related to operation
voltage, so Ipeak vs. operation voltage (Vop = Vbias +
Vout ) characteristics gives more complete information
(Fig. 15) [11]. This plot is qualitatively similar to the
base-grounded collector I-V curves of a bipolar transistor, though the measured Ipeak on the curve represents
a dynamic characteristic in the present case. As seen
in the gure, the output intensity (Ipeak ) is nearly constant while Vop is negative, indicating that the operation point is in the active region. When Vop moves to
forward diode voltages, Ipeak sharply drops with Vop ,
indicating the dynamic saturation. The wide voltage
swing obtained here is essential for ecient large signal
operation. Series resistance RS of a device can shift

Fig. 16 Photoresponse at various bias voltages measured for a


UTC-PD with WA = 80 nm and WC = 400 nm.

the Ipeak -Vop curve to the negative Vop side, which results in a narrower active region. However, this eect
was negligible in the fabricated device because RS was
only a few Ohms. The maximum peak current obtained
was 184 mA (averaged 460 kA/cm2 ) with an FWHM of
4.8 ps (Vout = 4.6 V on a 25- load).
6.5 On the Bias Dependence of Electron Velocity in
InP
Electron velocity overshoot plays an essential role in
UTC-PD operation, especially in reducing the space
charge in the collection layer and elevating saturation
output. Because an input optical pulse is very short
(subpicosecond), EOS measurements on UTC-PDs can
give more direct information for the overshoot than
microwave measurements on InP/InGaAs/InP doubleheterostructure bipolar transistors. Figure 16 represents the pulse photoresponse curves at various bias
voltages measured for a UTC-PD with WA = 80 nm
and WC = 400 nm [16]. The optical input level was

IEICE TRANS. ELECTRON., VOL.E83C, NO.6 JUNE 2000

946

made low (0.13 pJ/pulse) to maintain a small signal


condition (output current density of 10 kA/cm2 ).
With increasing the negative bias (Vbias ), the output signal rst shows a slight narrowing, then broadens signicantly. At Vbias = 10 V (electric eld E =
250 kV/cm), the waveform has trapezoid shape, clearly
indicating velocity saturation. From the half width of
7 ps, the drift velocity is evaluated at 6106 cm/s. The
output is the sharpest, F W HM = 2.5 ps (2.2 ps after
deconvolution), around Vbias of 0.5 to 1.0 V, where
the corresponding total band-bending (eld) is 1.25
1.75 V (31 to 44 kV/cm). This result is consistent with
the optimum bias for the overshoot where the potential
drop is about 2EL /q (EL is the energy separation between - and L-valley minima). Assuming a
Gaussian prole for the output waveform, FWHM of
2.2 ps (at Vbias = 1.0 V) yields an f3dB of 142 GHz
and a total delay time d (= 1/2f3dB ) of 1.1 ps. Although the absorption and collection layer travel times
were not resolved from the total delay time, the electron
velocity in the collection layer is estimated to be on the
order of 2 107 cm/s. In a dierent UTC-PD structure
(WA = 86 nm and WC = 230 nm), the output FWHM
of 1.44 ps was obtained with an f3dB of 235 GHz; a low
optimum bias around 0.75 V was also observed [17].
7.

Applications of InP/InGaAs UTC-PDs

The features of UTC operation mode can be summarized as: high current density, broad bandwidth, and
low operation voltage. In conguring a high bit-rate
broadband photoreceiver without electrical ampliers,
for example, a basic requirement is a short enough CR
time constant (determined by junction capacitance and
load resistance). UTC-PDs permit us to use a smaller
junction size for generating a certain output signal level
(1 Vpp /20 mApp for driving digital circuits), which
means that the possible bandwidth can be expanded
(e.g. Ipp = 30 mApp at f3dB = 174 GHz, [17]). Because
the operation voltage can be low, close to 0 V, a large
voltage swing is allowed even at a low bias voltage.
Furthermore, in order to achieve stable and reliable operation, a low power dissipation (a low junction temperature) is needed. Capability of the low operating
voltages can also answer this requirement.
A UTC-PD module has been used to directly drive
a D-type ip-op circuit as the receiver of a 40-Gbit/s
time-division multiplexing system [18][20]. Eye diagrams of receiver outputs are shown in Fig. 17 for
40 and 80-Gbit/s signals [18], [21]. Compared to the
case of post amplier utilization [22], [23], the input
sensitivity has been much improved. Millimeter-wave
generation with a UTC-PD has been demonstrated
at 40 and 60 GHz [24], [25]. Optical signals produced
by mode-locked laser diodes have been eciently converted to millimeter-wave signals. The highest power
was 12 dBm at 60 GHz with a photocurrent of 30 mA.

Fig. 17

Output signals of UTC-PD photoreceivers.

The bandwidth exceeded 60 GHz, so good conversion


eciency was obtained. The eld intensity of the radiation measured by EOS increased almost linearly with
photocurrents up to 30 mA. It has been also demonstrated that the UTC-PD has a good linearity range at
20 GHz even at zero bias voltage [11].
UTC-PDs have been monolithically integrated
with resonant tunneling diodes (RTDs) to realize new
types of high-speed RTD circuits with optical inputs.
The PD is used to receive the optical signal and switch
the circuit connected in parallel to the driver RTD.
Because the operating voltage of a high-speed RTD
is lower than 1 V, UTC-PDs are suitable as such integrated circuits. Optical-input RTD demultiplexers
and D-FFs have been demonstrated at RZ data rates
of up to 80 Gbit/s [26][28]. A similar RTD/UTC-PD
pair combination has been applied to an optoelectronic
clock recovery circuit, and clock signals were extracted
from 46.2 GHz optical data [29], [30].
Another important application of UTC-PDs is the
optical gates that consist of electroabsorption optical
modulators (EAMs) and high-power PDs. Using the
high output of the UTC-PD, the EAM can be driven
over a wide range of transmittance. The transmission
gates function as optical MUX, DEMUX, and wavelength convertors [31], [32]. In a 1:4 DEMUX experiment, the good sensitivity of 32.7 dBm at a bit-error
rate of 109 was achieved for 40-Gbit/s signals [31].
On the other hand, an invertor gate (using the opposite polarity of the PD output) works as a reshaping/regenerating repeater [33]. Due to the nonlinear
transmission characteristics of the EAM as regards input voltage, the ASE noise in the optical signal is signicantly reduced by the invertor function. Up to the
present, the maximum reported operation speed of a
PD-EAM gate was 40-Gbit/s [31]. Here, CR time constant of EAM and device packaging can limit operation

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

947

speeds. Traveling-wave design for an EAM and the


monolithic integration of two kinds of devices should
signicantly enhance the operation speed.

[2]

8.

[3]

Conclusion

The uni-traveling-carrier photodiode (UTC-PD) was


proposed as a new type of photodiode for the purpose
of increasing saturation output, motivated by the demand for superior photoreceivers combined with highgain, high-output optical ampliers. Prime features of
UTC-PDs are faster response, higher operation current,
and lower operation voltage. They originate from the
operation mode in which only electrons are used as active carriers. The velocity overshoot of electrons plays
a signicant role in reducing the space charge in the
depleted carrier collection layer. The small signal analysis introduced here can be used to predict the impact
of device parameters on bandwidth.
Pulse photoresponse measurements revealed how
the saturation behaviors of UTC-PD and pin-PD differ. Due to the current tail associated with hole carrier
storage, the bandwidth of a pin-PD rapidly decreases
with input. On the other hand, a UTC-PD oers a
much larger range of linearity. Enhancement of bandwidth with increasing input is unique to the UTC-PD.
The self-induced eld in the absorption layer accounts
for this eect. High output operation has been systematically investigated. In a typical UTC-PD (f3dB =
6090 GHz), good linearity is obtained at pulse peak
currents of up to 80 mApp , and the saturation level exceeds 180 mApp (J460 kA/cm2 , Vout = 4.6 Vpp ). Capability of low voltage operation contributes not only
to yielding a wider voltage swing but also to reducing
junction temperature.
Applications of UTC-PDs have already been
demonstrated.
They include photoreceivers without electronic ampliers (80-Gbit/s), millimeter-wave
generators (12 dBm at 60 GHz), PD/RTD optoelectronic integrated circuits (80-Gbit/s), and drivers for
optical modulators (40-Gbit/s). UTC-PDs will be
valuable in realizing optoelectronic/optomicrowave interfaces and drivers where current electronic technologies face diculties.

[4]

[5]

[6]

[7]

[8]

[9]

[10]

[11]

[12]

[13]

[14]

[15]

Acknowledgement
The authors would like to thank N. Watanabe for epitaxial growth, Y. Sato for his assistance in device fabrication, and Y. Matsuoka, S. Mitachi, K. Yamazaki,
Y. Ishii, H. Toba and H. Kyuragi for their continuous
support.

[16]

[17]

References
[18]
[1] K. Hagimoto, Y. Miyamoto, T. Kataoka, H. Ichino, and
O. Nakajima, Twenty-Gbit/s signal transmission using

simple high-sensitivity optical receiver, OFC 92 Tech. Dig.


p.48, 1992.
K.J. Williams, R.D. Esmann, and M. Dagenias, Nonlinearities in p-i-n microwave photodetectors, J. Lightwave
Technol., vol.14, pp.8496, 1996.
K.J. Williams and R.D. Esmann, Large-signal compression-current measurements in high-power microwave pin
photodiodes, Electron. Lett., vol.35, pp.8284, 1999.
H.F. Tayler, O. Eknoyan, C.S. Park, K.N. Choi, and
K. Chang, Traveling wave photodetectors, SPIE Proc.
1217, Optoelectronic Signal Processing for Phased-Array
Antennas II, pp.5963, 1990.
L.Y. Lin, M.C. Wu, T. Ito, T.A. Vang, R.E. Muller,
D.L. Sivco, and A.Y
Cho, High-power high-speed
photodetectorsDesign, analysis, and experimental demonstration, IEEE Trans. Microwave Theory & Tech.,
vol.45, pp.13201331, 1997.
T. Ishibashi, N. Shimizu, S. Kodama, H. Ito, T. Nagatsuma,
and T. Furuta, Uni-traveling-carrier photodiodes, Tech.
Dig., Ultrafast Electronics and Optoelectronics, (1997 OSA
Spring Topical Meeting), pp.166168, 1997.
K. Kurishima, H. Nakajima, T. Kobayashi, Y. Matsuoka,
and T. Ishibashi, Fabrication and characterization of
high-performance InP/InGaAs double-heterojunction bipolar transistors, IEEE Trans. Electron Devices, vol.41,
pp.13191326, 1994.
T. Ishibashi, S. Kodama, N. Shimizu, and T. Furuta, Highspeed response of uni-traveling carrier photodiodes, Jpn.
J. Appl. Phys., vol.36, pp.62636268, 1997.
T. Ishibashi, H. Fushimi, T. Furuta, and H. Ito, Unitraveling-carrier photodiodes for electromagnetic wave generation, Proc. THz 99, pp.3639, 1999.
T. Furuta, H. Ito, and T. Ishibashi, Photocurrent dynamics of uni-traveling-carrier and conventional photodiodes,
Ext. Abs. 26th Int. Symp. on Compound Semiconductors,
Berlin, Th B1-3, 1999.
T. Ishibashi, H. Fushimi, H. Ito, and T. Furuta, High
power uni-traveling-carrier photodiodes, Tech. Dig. Microwave Photonics (MWP 99), pp.7578, 1999.
Y. Muramoto, K. Kato, M. Mitsuhara, O. Nakajima,
Y. Matsuoka, N. Shimizu, and T. Ishibashi, Uni-travelingcarrier waveguide photodiode, Int. Conf. on Indium Phosphide and Related Materials, pp.407410, 1998.
P.-L. Liu, K.J. Williams, M.Y. Frankel, and R.D. Esmann,
Saturation characteristics of fast photodetectors, IEEE
Trans. Microwave Theory & Tech., vol.47, pp.12971301,
1999.
N. Shimizu, N. Watanabe, T. Furuta, and T. Ishibashi,
InP/InGaAs uni-traveling-carrier photodiode with improved 3-dB bandwidth of over 150 GHz, IEEE Photon.
Technol. Lett., vol.10, pp.412414, 1997.
N. Shimizu, N. Watanabe, T. Furuta, and T. Ishibashi, Improved response of uni-traveling-carrier photodiodes by carrier injection, Jpn. J. Appl. Phys., vol.37, Part 1, pp.1424
1426, 1997.
T. Ishibashi, H. Ito, H. Fushimi, T. Furuta, K. Sano, and
N. Shimizu, Uni-traveling-carrier photodiode as an optoelectronic driver, Tech. Dig., Ultrafast Electronics and Optoelectronics (1999 OSA Spring Topical Meeting), pp.106
108, 1999.
H. Ito, T. Furuta, S. Kodama, N. Watanabe, and
T. Ishibashi, InP/InGaAs uni-traveling-carrier photodiode
with 220 GHz bandwidth, Electron. Lett., vol.35, no.18,
pp.15561557, 1999.
Y. Miyamoto, M. Yoneyama, K. Hagimoto, T. Ishibashi,
and N. Shimizu, 40 Gbit/s high sensitivity optical receiver
with uni-traveling-carrier photodiode acting as a decision

IEICE TRANS. ELECTRON., VOL.E83C, NO.6 JUNE 2000

948

IC driver, Electron. Lett., vol.34, pp.214215, 1997.


[19] M. Yoneyama, Y. Miyamoto, T. Otsuji, A. Hirano,
H. Kikuchi, T. Ishibashi, and H. Miyazawa, Fully electrical 40-Gbit/s TDM system prototype and its application
to 160-Gbit/s WDM transmission, Tech. Dig. OFC 99,
pp.128130/ThI6-1, 1999.
[20] Y. Miyamoto, M. Yoneyama, T. Otsuji, K. Yonenaga, and
N. Shimizu, 40-Gbit/s TDM transmission technologies
based on High-Speed ICs, IEEE J. Solid-State Circuits,
vol.34, pp.12461253, Sept. 1999.
[21] Y. Miyamoto, K. Yonenaga, A. Hirano, N. Shimizu,
M. Yoneyama, H. Takara, K. Noguchi, and K. Suzuki,
1.04-Tbit/s DWDM transmission experiment based on
alternate-polarization 80-Gbit/s OTDM signals, 24th
ECOC, vol.3, pp.5557, 1998.
[22] M. Yoneyama, A. Sano, K. Hagimoto, T. Otsuji, K. Murata,
Y. Imai, S. Yamaguchi, T. Enoki, and E. Sano, A 40Gbit/s optical receiver circuit using InAlAs/InGaAs HEMT
IC chip set, 1EEE MTT-S, pp.461464, 1997.
[23] R. Ohhira, Y. Amamiya, T. Niwa, N. Nagano, T. Takeuchi,
C. Kurioka, T. Chuzenji, and K. Fukuchi, A highsensitivity 40-Gbit/s optical receiver using packaged GaAs
HBT-ICs, OFC 99, pp.155156, 1998.
[24] T. Ohno, S. Fukushima, Y. Doi, Y. Muramoto, and Y.
Matsuoka, Application of uni-traveling-carrier waveguide
photodiodes in base stations of a millimeterwave ber-radio
systems, Tech. Dig. Microwave Photonics (MWP99),
pp.253259, 1999.
[25] T. Nagatsuma, N. Sahri, M. Yaita, T. Ishibashi, N. Shimizu,
and K. Sato, All optoelectronic generation and detection of
millimeter-wave signals, Tech. Dig. Microwave Photonics
(MWP 98), pp.58, 1998.
[26] K. Sano, K. Murata, T. Akeyoshi, N. Shimizu, T. Otsuji,
M. Yamamoto, T. Ishibashi, and E. Sano, An ultra-fast
optoelectronic circuit using resonant tunneling diodes and
a uni-traveling-carrier photodiode, Electron. Lett., vol.34,
pp.215217, 1998.
[27] T. Akeyoshi, N. Shimizu, J. Osaka, M. Yamamoto,
T. Ishibashi, K. Sano, K. Murata, and E. Sano, An optoelectronic logic gate monolithically integrating resonant
tunneling diodes and a uni-traveling-carrier photodiode,
Jpn, J. Appl. Phys., vol.38, pp.12231226, 1999.
[28] K. Sano, K. Murata, T. Otsuji, T. Akeyoshi, N. Shimizu,
and E. Sano, 80 Gbit/s optoelectronic delayed ip-op
circuit using resonant tunneling diodes and uni-travelingcarrier photodiode, Electron. Lett., vol.35, pp.13761377,
1999.
[29] K. Murata, K. Sano, T. Akeyoshi, N. Shimizu, E. Sano,
M. Yamamoto, and T. Ishibashi, Optoelectronic clock
recovery circuit using resonant tunneling diode and
uni-traveling-carrier photodiode, Electron. Lett., vol.34,
pp.14241425, 1998.
[30] K. Murata, K. Sano, T. Akeyoshi, N. Shimizu, E. Sano,
M. Yamamoto, and T. Ishibashi, An optoelectronic clock
recovery circuit using a resonant tunneling diode and a
uni-traveling-carrier photodiode, IEICE Trans. Commun.,
vol.E82-B, no.8, pp.12281235, Aug. 1999.
[31] M. Yoneyama, Y. Miyamoto, K. Hagimoto, N. Shimizu,
T. Ishibashi, and K. Wakita, 40 Gbit/s optical gate using
optical modulator driven by uni-traveling-carrier photodiode, Electron. Lett., vol.34, pp.16071609, 1998.
[32] M. Yoneyama, Y. Miyamoto, K. Hagimoto, N. Shimizu,
T. Ishibashi, and K. Wakita, Simple wavelength converter using an optical modulator directly driven by a
uni-traveling-carrier photodiode, Electron. Lett., vol.34,
pp.22442245, 1998.
[33] Y. Kisaka, A. Hirano, M. Yoneyama, and N. Shimizu, Sim-

ple 2R repeater based on EA modulator directly driven by


uni-traveling-carrier photodiode, Electron. Lett., vol.35,
pp.10161017, 1999.

Tadao Ishibashi
received the B.S.,
M.S. and Ph.D. degrees in applied physics
from Hokkaido University in 1971, 1973,
and 1986, respectively.
Since joining
NTT Laboratories, Musashino, Tokyo, in
1973, he has been involved in research
of semiconductor devices and related material processing.
His work included
submillimeter-wave Si IMPATT diode oscillators, LPE growth of InP/InGaAs materials and their application to eld effect transistors, MBE growth of MQW laser diodes, and GaAsbased/InP-based heterostructure bipolar transistor ICs. He is
currently working on ultrahigh-speed optoelectronic devices and
their integration. During 1991 to 1992, he stayed at Max-PlanckInstitute, Stuttgart, as a visiting scientist. Dr. Ishibashi is a
member of the Japan Society of Applied Physics, and the Institute of Electrical and Electronics Engineers (IEEE). He received
the Ichimura Award in 1992 for the development of ballistic collection transistors.

Tomofumi Furuta
was born in Tokyo, Japan, on March 10, 1958. He received the B.S. and M.S. degrees in electrical engineering from Tokyo University
of Agriculture and Technology, Japan,
in 1981 and 1983, respectively, and the
Ph.D. degree from the University of Tokyo in 1986. In 1986, he joined the NTT
LSI laboratories, Kanagawa, Japan. He
has been engaged in the research of semiconductor physics and devices. Dr. Furuta is a member of The Japan Society of Applied Physics and
The American Institute of Physics.

ISHIBASHI et al.: InP/InGaAs UNI-TRAVELING-CARRIER PHOTODIODES

949

Hiroshi Fushimi
was born in Tokyo, Japan, on May 27, 1964. He received B.S. and the M.S. degrees in electrical engineering from the Nagaoka University of Technology, Nagaoka, Niigata,
Japan, in 1987 and 1989, and the Ph.D.
degree in electrical engineering from Keio
University, Yokohama, Kanagawa, Japan,
in 1998, respectively. He joined NTT
Transmission Systems Laboratories, Nippon Telegraph and Telephone Corporation, Yokosuka, Kanagawa, Japan, in 1989, where he was engaged
in research on high-speed GaAs ICs and coherent lightwave transmission systems. In 1993, he moved to NTT LSI Laboratories,
Atsugi, Kanagawa, Japan, where he has been engaged in research
on defect physics and electronically enhanced defect reactions in
semiconductors. He works now at NTT Photonics Laboratories,
Atsugi, Kanagawa, Japan. Dr. Fushimi is a member of the Japan
Society of Applied Physics, the Surface Science Society of Japan,
and the Japanese Association of Crystal Growth.

Satoshi Kodama
was born in Ohtsu,
Japan, in 1968. He received the B.S. degree in information engineering, and M.S.
and Ph.D. degrees in electrical engineering from Hokkaido University, Sapporo,
Japan, in 1991, 1993, and 1996, respectively. In 1996, he joined LSI laboratories,
Nippon Telegraph and Telephone Co., Atsugi, Japan. He is currently a Research
Engineer at Photonics Laboratories, Nippon Telegraph and Telephone Co., Atsugi, Japan. His research interest is integrated advanced optelectronics devices. Dr. Kodama is a member of the Japan Society of Applied Physics.

Hiroshi Ito
received the B.S. and
M.S. degrees in physics, and Ph.D. degree in electrical engineering, from Hokkaido University, Japan, in 1980, 1982,
and 1987, respectively. Since joining NTT
Laboratories in 1982, he has been involved
in research on growth and characterization of III-V compound semiconductors
using MBE and MOCVD, and their applications to devices such as heterojunction bipolar transistors (HBTs), eld effect transistors, lasers, and photodiodes. From 1991 to 1992, he
was at Stanford University as a visiting scientist. His current
research interests focus on ultrafast photonic devices and III-V
materials. Dr. Ito is a member of IEEE, the Physical Society of
Japan, and the Japan Society of Applied Physics.

Tadao Nagatsuma
received the
B.S., M.S. and Ph.D. degrees in electronic engineering from Kyushu University, Fukuoka, Japan, in 1981, 1983, and
1986, respectively. He joined the Nippon Telegraph and Telephone Corporation (NTT), Atsugi Electrical Communications Laboratories, Kanagawa, Japan,
in 1986, and currently is a distinguished
technical member, senior research scientist, supervisor at NTT Telecommunications Energy Laboratories. His current research involves highspeed electronics and millimeter-wave photonics, and their application to sensors and communications. Dr. Nagatsuma is a member of the Optical Society of America, the IEEE, and the Japan
Society of Applied Physics. He was the recipient of the 1989
Young Engineers Award from IEICE, the 1992 Andrew R. Chi
Best-Paper Award from IEEE, the 1997 Okochi Memorial Award,
and the 1998 Japan Microwave Prize.

Naofumi Shimizu
was born in
Osaka, Japan, in September 1962. He received the B.E. and M.E. degrees in engineering physics from Kyoto University,
Kyoto, Japan, in 1986 and 1988, respectively. In 1988, he joined NTT LSI Laboratories, Kanagawa, Japan. He was engaged in research and development on IIIV high-speed devices. Since 1998, he has
been with Network Innovation Laboratories, where he has been engaged in research on high-speed lightwave transport systems.

Yutaka Miyamoto
was born in Tokyo, Japan, on December 8, 1963. He
received the B.E. and M.E. degrees in
electrical engineering from Waseda University, Tokyo, Japan, in 1986 and 1988,
respectively.
In 1988, he joined the
NTT Transmission Systems Laboratories,
Yokosuka, Japan, where he engaged in
research and development on high-speed
optical communications systems including
the 10-Gbit/s terrestrial optical transmission system called FA-10G system. He is currently a senior research engineer of NTT Network Innovation Laboratories, Yokosuka, Kanagawa, Japan. His current research interest includes
ultra-high-speed communication systems and their related devices. Mr. Miyamoto is a member of the Institute of Electrical
and Electronics Engineers (IEEE). He received the best paper
award of the rst optoelectronics and communication conference
(OECC96) in 1996 from the IEICE.

S-ar putea să vă placă și