Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

IET Optoelectronics

Special Issue: Selected Papers from Semiconductor and Integrated


Optoelectronics (SIOE) 2017

Physical modelling and experimental ISSN 1751-8768


Received on 13th June 2017
Revised 14th August 2017
characterisation of InAlAs/InGaAs avalanche Accepted on 11th September 2017
E-First on 10th October 2017
photodiode for 10 Gb/s data rates and higher doi: 10.1049/iet-opt.2017.0068
www.ietdl.org

Omar S. Abdulwahid1, James Sexton1, Ioannis Kostakis2, Kawa Ian2, Mohamed Missous1
1School of Electrical and Electronic Engineering, University of Manchester, Manchester, UK
2Integrated Compound Semiconductors, Manchester, UK
E-mail: m.missous@manchester.ac.uk

Abstract: InAlAs/InGaAs avalanche photodiodes (APD) were simulated using physical device models, then designed and
fabricated to detect light in the wavelength range from 1.3 to 1.55 µm. DC characterisation under dark and light conditions were
performed at room temperature to measure and investigate the performance of the APD. High-frequency characterisation was
carried out on the device to extract the diode intrinsic and extrinsic parameters. The work reported here focuses on the dark and
light physical device simulations (both under DC and AC conditions) which were accomplished using Atlas SILVACO tool. The
effect of electron velocity overshoot was considered for accurate bandwidth modelling. All measured data are in excellent
agreement with the modelled ones. The internal device gain of the APD is 45 at −23.5 V leading to a ∼220 GHz gain–bandwidth
product. This successful APD model can be exploited to further improve the diode structure for higher data rate applications
beyond 10 Gb/s.

1 Introduction circuit model and SILVACO physical model, showed high


correlation with the measured data which validates the models
The avalanche photodiode (APD) is the preferred photodiode used. Specifically, the junction capacitance at different bias
receiver device in telecommunication detection systems due to its voltages was extracted from both models and the obtained values
internal gain which leads to high sensitivity at low optical power. show excellent agreements with the measured ones.
However, the random nature of the impact ionisation process in the The structure reported in this work was grown using molecular
APD generally introduces excess noise which depends on exact beam epitaxy on a 620 µm thick semi-insulating InP substrate as
semiconductor materials used in the device [1]. In practice, the shown in Fig. 1. It comprises p-type In0.53Ga0.47As top and n-type
APD is always used in tandem with a trans-impedance amplifier
In0.53Ga0.47As bottom contacts, p-type In0.53Ga0.47As top and n-
can be incorporated to adjust the gain, improve the signal-to-noise
ratio and the bandwidth of the optical receiver system [2]. APD type In0.52Al0.48As bottom cladding layers, undoped In0.53Ga0.47As
based on InP and InAlAs multiplication layers have been absorber layer, p-type In0.52Al0.22Ga0.25As grading layer, p-type
extensively studied by many groups [1, 3–6]. The InAlAs material In0.52Al0.48As charge sheet layer, and undoped In0.52Al0.48As
is an electron multiplication material with a k-ratio of 0.29–0.5 [7], multiplication layer. The top and bottom contacts are heavily doped
while InP is a hole multiplication material with a k-ratio of 0.4–0.5 (>2 × 1019 cm−3) and 1 × 1019 cm−3 with thicknesses of ∼100 and
[8]. Moreover, an APD with InAlAs multiplication layer has better 500 nm, respectively. The highly doped contacts help to reduce the
stability compared with the one based on InP multiplication layer series resistance (RS) which leads to improvements in the
[3]. The lowest reported k-ratio material is silicon with k = 0.03–0.1 frequency response of the device. The absorber is relatively thick
[9]. However, silicon is not lattice matched with InGaAs material, (∼1.5 µm), but this makes the APD efficient at absorbing light in
which limits its use for 1.3–1.55 wavelength telecommunication the 1.3–1.6 µm wavelength region. The grading layer thickness is
applications, even though attempts at using mismatched Si–Ge are 50 nm which improves the frequency response of the APD by
underway [10]. reducing the band discontinuity at the interface with the charge
The work reported here can be summarised as follow. Firstly, an
layer. The charge layer has a doping profile of ∼1 × 1018 cm−3 with
InGaAs–InAlAs APD with a light window size of 30 µm was
a thickness of 50 nm. The main function of this layer is adjusting
designed and then experimentally fabricated and characterised in
the electric field of the device. Finally, the multiplication layer with
terms of its DC and high-frequency optical characteristics.
a thickness of 200 nm is buried under the charge layer.
Secondly, high-frequency equivalent circuits from fabricated
devices were built up to 40 GHz to extract key diode parameters,
including junction capacitance, series resistance, and junction 2 Experimental and simulation characterisation
resistance utilising Advanced Design System (ADS) software from details
Keysight Technologies. Extracting the APD parameters is
The APD was experimentally characterised to determine its
necessary to calculate its cut-off frequency and predict its
electrical and optical characteristics. The electrical characterisation
performance for high-frequency applications. Finally, physical
was performed under dark conditions and at room temperature to
models of the APD structure, using the Atlas SILVACO simulation
measure the dark current, capacitance–voltage (C–V)
tool, including dark current, photocurrent current, C–V
characteristics, and high-frequency s-parameter measurements up
characteristics, optical 3 dB bandwidth, and gain, are developed
to 40 GHz at different bias voltages. An Anritsu VNA was used to
and validated by an experimental APD epi-layer structure. To the
collect the s-parameter data as well as measuring the C–V
best of the authors knowledge, this is the first full virtual wafer
characteristics. For the optical characterisation, a 1.55 µm
fabrication physical modelling (DC and C–V characteristics, and
wavelength laser with −30 dBm output power (1 µW) was utilised
optical 3 dB bandwidth) of an InGaAs–InAlAs APD in SILVACO
to illuminate the device. The laser diameter was 5 µm. A
using the concept of electron velocity overshoot. The simulated
Lightwave Component Analyser (HP 8703A) was employed to
results for both models, i.e. high-frequency small-signal equivalent
measure the 3 dB optical bandwidth of the device. The bias voltage

IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10 5


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)
Fig. 3  Measured and simulated s-parameters Smith charts of the open and
short APD structures and corresponding equivalent circuits

and extrinsic component of the device such as series resistance


(RS), junction resistance (RJ), junction capacitance (CJ), pad
capacitance (CP), and pad inductance (LP). The parasitic
components are introduced by the GSG coplanar waveguide
Fig. 1  APD epi-layer structure
configuration. The optimisation of the CPW dimensions is required
to further reduce the parasitic capacitance caused by fringing effect
without increase in conductor losses or change in the designated
characteristic impedance (50 Ω) [15–17]. High-frequency s-
parameter measurements were performed for the open, short, and
actual structures from 40 MHz to 40 GHz at different bias voltages.
Fig. 2 depicts the GSG coplanar waveguide one-port structure of
the APD.
The dimension of the GSG coplanar waveguide was optimised
to give an impedance of 50 Ω. All measurements were performed
in the dark at room temperature. As there is no standard model for
the APD in ADS, an equivalent circuit model was built and the
fitting was made with measured data to validate this model and
extract the intrinsic and extrinsic components of the InAlAs/
InGaAs APD. The equivalent circuit for the open structure was
built in ADS and is represented by a capacitor only (CP), while the
short structure is represented by an inductor (LP). The simulated s-
parameters of the equivalent circuits were fitted with the measured
ones to extract CP and LP. The measured and simulated s-
Fig. 2  GSG coplanar waveguide structure of the APD parameters represented on a Smith chart for the open and short
devices are shown in Fig. 3. The excellent agreement between the
is applied using a DC supply HP4142B connected to the Analyser measured and simulated data validates the equivalent circuits used
through a bias-T. to extract CP and LP. The extracted CP and LP are 8 fF and 40 pH,
The main objective of this work was to build a quantitative and respectively. The small parasitic capacitance (CP) comes from the
predictive physical model for the APD photo-detector to validate optimised coplanar waveguide design process. In the same manner,
the measured electrical and optical characteristics and which can the equivalent circuit for the actual structure was built at different
then be used for further device improvements. Therefore, bias voltages and its s-parameters were compared with the
numerical simulations of the InAlAs/InGaAs APD photodetector measured ones to extract the intrinsic components (RS, CJ, and RJ)
under dark and light conditions and at room temperature were of the device as shown in Fig. 4. CP and LP were de-embedded
carried out using the Atlas SILVACO tool. Device structure and from the total equivalent circuit. Fig. 4 depicts the measured and
type determines the required physical models to use. To model the equivalent circuit s-parameters of the device at −15 V.
impact ionisation process of the APD, an IMPACT SELBER model The equivalent circuit s-parameters agree extremely well with
was used as will be discussed later in Section 4. As the SILVACO the measured ones over the frequency range 40 MHz–40 GHz with
library does not contain material parameters for InGaAs, InAlAs, simulated and measured curves being essentially identical (Figs. 3
and InAlGaAs, all III–V material parameters were obtained both and 4). Table 1 lists the extracted component at different bias
from the literatures and from validation from a number of devices voltages.
studied over the years in our lab [4, 11–14]. The fully depleted junction capacitance of the APD is 162 fF.
The small RS (10 Ω) of the APD is due to the highly doped profile
3 S-parameter measurement and small-signal of the top and bottom contact layers. The series resistance (RS) is
radio frequency equivalent circuit extraction of the mainly due to contributions from the top contact resistance and
APD spreading and epi-layer resistances [18]. Reducing the separation
between the top and bottom contact also helps minimise the series
The device was fabricated into a circular mesa form with an resistance of the APD. When the reverse bias voltage increased
aperture window of 30 µm as shown in Fig. 2. High-frequency beyond punch-through, a voltage drop occurs through the series
characterisation plays an important role in determining the intrinsic and load resistances. At higher gain levels, a large photocurrent

6 IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)
Fig. 6  InAlAs/InGaAs APD measured and simulated dark currents
Fig. 4  Measured and simulated s-parameters Smith charts and equivalent
circuit of the APD at −15 V bias voltage Table 2 Key fitting parameters used in SILVACO physical
modelling
Table 1 InAlAs/InGaAs APD extracted parameters
Parameter Absorber Multiplication Grading layer
Component Bias = −15 Bias = −18 V Bias = −21 V layer layer (InAlAs) (InAlGaAs)
CJ, fF 162 158 159 (InGaAs)
RS, Ω 10 9.6 9.6 electron mobility, 11,000 4500 2300
RJ, kΩ 15 15 15 V/cm2
CP, fF 8 8 8 energy gap, eV 0.75 1.44 ∼0.99
LP, pH 40 40 40 affinity, eV ∼4.5 ∼4.25 ∼4.38
permittivity 13.9 12.2 ∼12.5
electron carrier life 100 — —
time, ns
electron effective ∼0.042 ∼0.085 ∼0.06
mass
hole effective mass ∼0.46 ∼0.6 ∼0.61

used. This model was then used to simulate the C–V


characteristics, photocurrent, and optical 3 dB bandwidth. In
SILVACO, a bias voltage was applied as shown in Fig. 5 and the
device simulated to extract the dark current under the same
conditions. Fig. 6 shows the measured and simulated dark currents
of the InAlAs/InGaAs APD. The modelling process of the APD
photodetector was performed using two assumptions. The first
Fig. 5  Three-dimensional and 2D sectional views of the InAlAs/InGaAs assumption was based on the dark current characterisation, where
APD in SILVACO no electron–hole generation is defined in the absorber layer, and
where the multiplication region has a low electric field distribution.
flows in the series resistance resulting in undesired behaviour Therefore, electrons travel with their normal velocity in both
leading to a non-linear relationship between output current and multiplication and charge sheet layers. According to that, the
applied light [19]. The intrinsic cut-off frequency calculated using electron velocity was set to 2.5 × 106 and 1 × 107 cm/s in the charge
the usual expression 1/2πRSCJ is 105 GHz and the extrinsic cut-off and multiplication layers, respectively. The electron velocity in the
frequency is 16 GHz when used in a 50 Ω load as would be the absorption layer was set to 1.5 × 107 cm/s. The key fitting
case for an APD. Extracting the junction capacitance of the APD parameters used in SILVACO modelling are shown in Table 2.
using the high-frequency small-signal equivalent circuit model is The electron mobilities of each layer (InGaAs, InAlAs, and
necessary to validate the SILVACO physical model which exhibits InAlGaAs) are different depending upon the doping profile of each
almost the same value when the APD is fully depleted as will be layer. All required values were obtained from the literatures in [12–
discussed in Section 4. 14]. For the simulation of the dark current, several models have to
be considered. The Shockley–Read–Hall (SRH) model and Fermi–
4 Experimental and physical modelling results Dirac statistics were used to model the generation-recombination
and carrier drift-diffusion processes both of which make large
The APD was modelled using the same dimensions as the contributions to the total dark current. The basic analytical
fabricated structure such as aperture window size, mesa size, anode expression of these models can be found in details in [20]. The
and cathode diameter sizes, anode to cathode separation, and band-to-band tunnelling current was not considered due to the
region thicknesses. Fig. 5 depicts the three-dimensional (3D) APD inclusion of the graded and charge sheet layers which provide
configuration and its 2D sectional view in the Atlas SILVACO tool. enough electric field separation between the absorber and
multiplication layers. Band-to-band tunnelling current starts to
The dark current of the device was measured up to −25 V dominate the dark current of the APD when the multiplication
reverse bias voltage using a probe station under dark room region is smaller than 100 nm [21]. The most important phenomena
conditions and at room temperature. The fitting process of the dark to take in account for accurate physical simulation of the dark
current is crucial to validate the model and material parameters

IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10 7


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)
measured and simulated junction capacitances at different reverse
bias voltages. The doping profile of the charge sheet layer is a
critical factor in fitting the simulated junction capacitance to the
measured one. The fitting process indicated that the doping of this
layer is ∼6.2 × 1017 cm−3. The simulated junction capacitance is in
an excellent agreement with the measured one except for a
difference of ∼12% from −18 to −21 V bias voltages. This
difference could be due to the wider simulated depletion region in
the SILVACO tool, resulting in a smaller capacitance value.
The device has a wide margin voltage range where the punch-
through voltage (VPT) occurs at (−12.5 V), which is far enough
from the breakdown voltage (VBR = −24 V). At the punch-through
voltage (VPT), the junction capacitance value starts to drop due to
the expansion of the depletion region. The fully depleted junction
capacitance is 160 fF which matches the extracted value from the
high-frequency small-signal equivalent circuit. The junction
capacitance can be further minimised by enlarging the absorber
thickness, but this would lead to increasing the carrier transient
Fig. 7  Measured and simulated C–V characteristic of the InAlAs/InGaAs time and as a result degrading the bandwidth.
APD The second assumption in the modelling process was based on
the light characterisation which takes into account the generation
current is the impact ionisation process which is described by the process of the electron–hole pair in the absorber layer and the high
following equation [20]: electric field distribution in the charge and multiplication layers.
The optical generation rate is calculated in SILVACO using the
G = αn J n + αp J p (1) formula [20]:

where G is the generation rate of the electron–hole pairs, αn and αp P ∗ λ − αy


G = η0 αe (4)
are the electron and hole impact ionisation coefficients, and Jn and hc
Jp the electron and hole current densities. In SILVACO, FLDMOB,
where η0 is the material quantum efficiency, P* represents the
and a local field, IMPACT SELBER models were used to model
effect of absorption losses and transmission and reflection factors,
the electric field mobility dependency and impact ionisation rate of
λ is the light wavelength, α the material absorption coefficient, h
electron and hole in the multiplication region, respectively. Both
the Planck constant, and c the speed of light. The photo-response
models are necessary to fit the dark current and breakdown voltage
characteristics of the InAlAs/InGaAs were measured and modelled.
(VBR). The IMPACT SELBER model which is a variation of the
A laser light was utilised with a wavelength of 1.55 µm to generate
classical Chynoweth model was used to determine the ionisation electron–hole pairs in the absorption layer. The laser power was 1 
coefficients (αn and αp) using the following equations: µW as shown in Fig. 5. In the photocurrent simulation process, the
same models (SRH, Fermi–Dirac statistics, IMPAT SELBER, and
BETAN
BN FLDMOB) and fitting parameters of the dark current and C–V
αn = ANexp − (2)
E characteristics were used except that the electron velocity in the
absorption and charge sheet layers were set to 2 × 107 and 5 × 107 
BETAP cm/s, respectively, as the electric field is higher under light
BP
αp = APexp − (3) conditions. In [4], a Monte Carlo model was used to simulate the
E
optical characteristics of the APD. In [9], it was shown that in thin
where E is the electric field across the structure. AN, AP, BN, BP, multiplication regions and high electric fields, the electron can
BETAN, and BETAP are the impact statement parameters [20]. travel with a speed that is much higher than its saturation velocity.
Through the simulation, SILVACO calculates these parameters This was also confirmed in [22], where the carrier velocity used in
according to the material parameters of the charge sheet the model was much higher than the saturation velocity for an APD
multiplication regions (lattice temperature and energy gap) as well with a 200 nm InAlAs multiplication region. This concept was
as the required model for the impact ionisation process. The further explored in our model where the charge sheet and
calculation process of the parameters can be found in details in multiplication layers have a high electric field profile as shown in
[20]. The output window of the SILVACO resulted in values of 8.6  Fig. 8. As can be clearly seen, the electric field is ∼690 and 700 
kV/cm at −18 and −20 V bias voltages, respectively. At higher
× 106 cm−1, 2.3 × 107 cm−1, 3.5 × 106 cm−1, 4.5 × 106 cm−1, 1, and
electric field, the newly generated electrons, due to the impact
1 for AN, AP, BN, BP, BETAN, and BETAP, respectively. From
ionisation process, tend to populate the Г valley where the
Fig. 6, it is clear that there is excellent fit between experimental
electrons have lighter effective mass. This then lead to a carrier
and simulated data which validates the physical models used. The
velocity overshot behaviour. The measured and simulated dark and
excellent agreement is due to the appropriate material parameters
photocurrents are shown in Fig. 9a. The appropriate material
and models used. The breakdown voltage is −24 V (defined at 0.1 
parameter and models used resulted in excellent agreement
mA dark current). In Silvaco, the doping profile and the thickness
between the measured and simulated results. Thereafter, this model
of the charge sheet layer are the key factors in adjusting the VBR of
was used to simulate the frequency response of the InGaAs/InAlAs
the device to be fitted with the measured one and this therefore APD by only changing the electron velocity in the charge sheet
allows for a determination of the actual doping and thickness of the layer according to the gain value. Under dark conditions, there is
specified layer for further improvements. The dark current for the no electron–hole pair generation and no injected electron into the
device is <14 nA at (90%VBR), which makes the device efficient multiplication region and as a result, the dark current does not
for high-sensitivity receivers. change much before and after the punch-through voltage (VPT). For
Similarly, the measurement of the device capacitance–voltage this reason, VPT cannot be determined from the dark current.
(C–V) characteristic is important in determining the actual doping
However, VPT can easily be determined from the photocurrent
profile and layers thicknesses of the fabricated device. The total
capacitance including the junction and parasitic capacitances was curve (Fig. 9a) and is around −12.5 V which matches well with the
measured at different bias. The parasitic capacitance comes from value obtained from the C–V characteristics. The photocurrent is
the GSG coplanar waveguide pad. The junction capacitance of the equal to the dark current for voltages < punch trough voltage (VPT).
device was then extracted as shown in Fig. 7 which depicts the For voltages >VPT, the carrier starts to be injected into the

8 IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)
Fig. 8  Simulated electric field distribution of the InAlAs/InGaAs APD

Fig. 10  Measured and simulated


(a) S21 response, (b) Bandwidth versus bias voltages of the InAlAs/InGaAs APD

injected primary photocurrent at −13.4 bias voltage as depicted in


Fig. 9b. The internal gain increases with increasing the bias voltage
after the punch-through voltage where it is clear that the APD can
provide a gain >10 at 90%VBR. High-sensitivity receivers require
the high gain that an APD provides. However, the excess noise
factor increases at high multiplication gain which in turn degrades
the signal-to-noise ratio. The 3 dB optical bandwidth of the device
was calculated from the measured S21 frequency response. S21
response represents the ratio of the measured current signal at the
GSG probe to the applied laser power. S21 frequency response was
measured at different bias voltages at 1 µW input laser power.
The simulation process of the S21 frequency response was
performed using the same model used to simulate the photocurrent.
The fitting parameter (charge sheet layer electron velocity) was
changed according to the gain values because of the field-velocity
dependency. The charge sheet layer electron velocity was adjusted
to fit the simulated optical bandwidth with the measured one at
different bias voltages according to the velocity overshot in the thin
multiplication layer. At gain values of 5–10, the electron velocity
in the absorption layer and multiplication layer were set to 2 × 107
Fig. 9  Measured and simulated and 1 × 107 cm/s, respectively. The effect of the velocity overshoot
(a) Total currents, (b) Multiplication gain of the InAlAs/InGaAs APD was applied on the electron velocity in the charge sheet layer.
Therefore, the electron velocity of the charge sheet layer was
multiplication region, which leads to an increase in the varied according to the applied bias voltage and it was set to 6 × 
photocurrent. At the breakdown voltage (VBR = −24 V), a large 107, 5.5 × 107, 5 × 107, and 4 × 107 cm/s at −18, −19, −20, and −21 
number of impact ionisation events occur resulting in a large V bias voltage, respectively. Fig. 10a shows the measured and
photocurrent exceeding 0.1 mA. The measured and simulated simulated S21 response at −20 V bias voltage and Fig. 10b shows
multiplication gain is given by the ratio of the photocurrent to the the variation of bandwidth with applied bias.

IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10 9


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)
The excellent fit between the measured and simulated results [4] Ma, F., Li, N., Campbell, J.C.: ‘Monte Carlo simulations of the bandwidth of
InAlAs avalanche photodiodes’, IEEE Trans. Electron Devices, 2003, 50, pp.
validates the SILVACO APD models developed which should then 2291–2294
provide further predictive behaviour of the device characteristics [5] Tan, L.J.J., Ong, D.S.G., Ng, J.S., et al.: ‘Temperature dependence of
under different epitaxial layers and device layout configurations for avalanche breakdown in InP and InAlAs’, IEEE J. Quantum Electron., 2010,
25 G/s applications and above. The device has an optical 46, pp. 1153–1157
[6] Haško, D., Kováč, J., Uherek, F., et al.: ‘Avalanche photodiode with sectional
bandwidth of 6.7 GHz at −20 V bias voltage, which can be used for InGaAsP/InP charge layer’, Microelectron. J., 2006, 37, pp. 483–486
10 G/s applications. At higher bias voltage close to the breakdown [7] Watanabe, I., Torikai, T., Makita, K., et al.: ‘Impact ionization rates in (100)
voltage, the electrons start to scatter from Г to L and X valleys Al (0.48) In (0.52) As’, IEEE Electron Device Lett., 1990, 11, p. 437
where the effective masses are higher. Therefore, the electrons [8] Campbell, J.C.: ‘Recent advances in telecommunications avalanche
photodiodes’, J. Lightwave Technol., 2007, 25, pp. 109–121
travel with lower speeds resulting in a reduced operating [9] Kang, Y., Liu, H.-D., Morse, M., et al.: ‘Monolithic germanium/silicon
bandwidth as can be seen in Fig. 10b. In our model at −22 V bias avalanche photodiodes with 340 GHz gain–bandwidth product’, Nat.
voltage, the electron velocity of the charge sheet layer was reduced Photonics, 2009, 3, pp. 59–63
to 1 × 107 cm/s. The measured gain-bandwidth of the device is 220  [10] Jang, K.-S., Kim, S., Kim, I.G., et al.: ‘High performance Ge-on-Si avalanche
photodetector’. SPIE OPTO, 2016, pp. 97531C–97531C-6
GHz at −23.5 V bias voltage (for a gain of 45). [11] Meier, H.T.J.: ‘Design, characterization and simulation of avalanche
photodiodes’ (ETH ZURICH, 2011)
[12] ‘In(1-x-y)Al(x)Ga(y)As, physical properties: datasheet from Landolt-
5 Conclusion Börnstein – Group III condensed matter’, in Madelung, O., Rössler, U.,
In this work, detailed physical modelling of an InAlAs/InGaAs Schulz, M. (Eds.): ‘Group IV elements, IV-IV and III-V Compounds. Part b –
electronic, transport, optical and other properties’, (Springer-Verlag, Berlin
avalanche photodetector including dark and light DC Heidelberg, 2002), http://dx.doi.org/10.1007/10832182_36
characteristics, frequency response, and C–V simulation using [13] ‘Al(x)In(1-x)As physical properties: datasheet from Landolt-Börnstein –
Atlas SILVACO were presented which agreed extremely well with Group III condensed matter’, in Madelung, O., Rössler, U., Schulz, M. (Eds.):
measured data. The intrinsic and extrinsic diode parameters were ‘Group IV elements, IV-IV and III-V compounds. Part b – electronic,
transport, optical and other properties’, (Springer-Verlag, Berlin Heidelberg,
accurately extracted up to 40 GHz using a small-signal equivalent 2002), http://dx.doi.org/10.1007/10832182_12
circuit technique. The velocity overshoot of electron gave an [14] ‘Ga(x)In(1-x)As, physical properties: datasheet from Landolt-Börnstein –
optimal modelled bandwidth which fits the measured data at Group III condensed matter’, in Madelung, O., Rössler, U., Schulz, M. (Eds.):
different bias voltages. This successful physical model provides ‘Group IV elements, IV-IV and III-V compounds. Part b – electronic,
transport, optical and other properties’, (Springer-Verlag, Berlin Heidelberg,
excellent quantitative predictions of the APD characteristics which 2002), http://dx.doi.org/10.1007/10832182_17
can be useful to further improve the device performances. This [15] Gopinath, A.: ‘Losses in coplanar waveguides’, IEEE Trans. Microw. Theory
model represents a platform to design APDs operating at high data Tech., 1982, 30, pp. 1101–1104
rates, e.g. 25 Gb/s receiver systems and higher. [16] Pucel, R.A.: ‘Design considerations for monolithic microwave circuits’, IEEE
Trans. Microw. Theory Tech., 1981, 29, pp. 513–534
[17] Frankel, M.Y., Gupta, S., Valdmanis, J.A., et al.: ‘Terahertz attenuation and
6 Acknowledgments dispersion characteristics of coplanar transmission lines’, IEEE Trans.
Microw. Theory Tech., 1991, 39, pp. 910–916
We are grateful for the support of the UK's Engineering and [18] Kanaya, H., Maekawa, T., Suzuki, S., et al.: ‘Structure dependence of
Physical Sciences Research Council under grant (EPSRC-EP/ oscillation characteristics of resonant-tunneling-diode terahertz oscillators
associated with intrinsic and extrinsic delay times’, Jpn. J. Appl. Phys., 2015,
P006973/1) ‘Future Compound Semiconductor Manufacturing 54, p. 094103
Hub’. [19] H. Photonics: ‘Characteristics and use of Si APD (avalanche photodiode)’,
Solid State Division, 2000
[20] I. Silvaco: ‘ATLAS user's manual device simulation software’, Santa Clara,
7 References CA, 2010
[21] Nakata, T., Takeuchi, T., Makita, K., et al.: ‘An ultra high speed waveguide
[1] Saleh, M.A., Hayat, M.M., Sotirelis, P.P., et al.: ‘Impact-ionization and noise
avalanche photodiode for 40-Gb/s optical receiver’. 27th European Conf. on
characteristics of thin III–V avalanche photodiodes’, IEEE Trans. Electron
Optical Communication, 2001. ECOC'01, 2001, pp. 564–565
Devices, 2001, 48, pp. 2722–2731
[22] Hambleton, P., Ng, B., Plimmer, S., et al.: ‘The effects of nonlocal impact
[2] Lau, K., Tan, C., Ng, B., et al.: ‘Excess noise measurement in avalanche
ionization on the speed of avalanche photodiodes’, IEEE Trans. Electron
photodiodes using a transimpedance amplifier front-end’, Meas. Sci. Technol.,
Devices, 2003, 50, pp. 347–351
2006, 17, p. 1941
[3] Meng, X., Tan, C.H., Dimler, S., et al.: ‘1550 nm ingaas/InAlAs single photon
avalanche diode at room temperature’, Opt. Express, 2014, 22, pp. 22608–
22615

10 IET Optoelectron., 2018, Vol. 12 Iss. 1, pp. 5-10


This is an open access article published by the IET under the Creative Commons Attribution License
(http://creativecommons.org/licenses/by/3.0/)

You might also like