Junction-less phototransistor with nanowire channels, a modeling study Anita Fadavi Roudsari,

advertisement
Junction-less phototransistor with nanowire
channels, a modeling study
Anita Fadavi Roudsari,1,* Simarjeet S. Saini,1 Nixon O,2 and M. P. Anantram3
1
Department of Electrical and Computer Engineering, University of Waterloo, 200 University Avenue W. Waterloo,
ON, Canada
2
Teledyne DALSA Inc., 605 McMurray Rd, Waterloo, ON, Canada
3
Department of Electrical Engineering, University of Washington, 185 Stevens Way, Paul Allen Center-Room
AE100R, Seattle, Washington, USA
*
afadavi@uwaterloo.ca
Abstract: We propose a new nanowire based, junction-less phototransistor,
that consists of a channel with both wide and narrow regions to ensure
efficient light absorption and low dark current, respectively. While the light
is absorbed in the wide region, the narrow region allows for ease of band
engineering. We also find that a nanowire in the source can further boost
the optical gain. The proposed device, which can potentially detect very low
light intensities, does not rely on complicated doping profiles, but instead
uses suitably designed gates. Our calculations show the detection of a
photon flux as low as 35 per second.
© 2014 Optical Society of America
OCIS codes: (040.5160) Photodetectors; (040.3780) Low light level; (040.6070) Solid state
detectors; (040.6040) Silicon; (230.5160) Photodetectors; (250.0040) Detectors.
References and links
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
D. Shiri, Y. Kong, A. Buin, and M. P. Anantram, “Strain Induced Change of Bandgap and Effective Mass in
Silicon Nanowires,” Appl. Phys. Lett. 93(7), 073114 (2008).
A. Zhang, S. You, C. Soci, Y. Liu, D. Wang, and Y. H. Lo, “Silicon Nanowire Detectors Showing
Phototransistive Gain,” Appl. Phys. Lett. 93(12), 121110 (2008).
R. Agarwal and C. M. Lieber, “Semiconductor Nanowires: Optics and Optoelectronics,” Appl. Phys., A Mater.
Sci. Process. 85(3), 209–215 (2006).
J. H. Park, S. H. Seo, I. S. Wang, H. J. Yoon, J. K. Shin, P. Choi, Y. C. Jo, and H. Kim, “Active Pixel Sensor
Using a 1x16 Nano-Wire Photodetector Array for Complementary Metal Oxide Semiconductor Imagers,” Jpn. J.
Appl. Phys. 43(4B), 2050–2053 (2004).
H. G. Choi, Y. S. Choi, Y. C. Jo, and H. Kim, “A Low-Power Silicon-on-Insulator Photodetector with a
Nanometer-Scale Wire for Highly Integrated Circuit,” Jpn. J. Appl. Phys. 43(6B), 3916–3918 (2004).
A. Kranti, R. Yan, C.-W. Lee, I. Ferain, R. Yu, N. Dehdashti Akhavan, P. Razavi, and J.-P. Colinge,
“Junctionless nanowire transistor (JNT): Properties and design guidelines,” in Proceedings of the European
Solid-State Device Research Conference (Sevilla, 2010), 357 −360.
S. Han, W. Jin, D. Zhang, T. Tang, C. Li, X. Liu, Z. Liu, B. Lei, and C. Zhou, “Photoconduction Studies on GaN
Nanowire Transistors under UV and Polarized UV Illumination,” Chem. Phys. Lett. 389(1-3), 176–180 (2004).
K. H. Kim, K. Keem, D. Y. Jeong, B. Min, K. Cho, H. Kim, B. M. Moon, T. Noh, J. Park, M. Suh, and S. Kim,
“Photocurrent of Undoped, n- and p-Type Si Nanowires Synthesized by Thermal Chemical Vapor Deposition,”
Jpn. J. Appl. Phys. 45(5A 5R), 4265–4269 (2006).
W. Kim and K. S. Chu, “ZnO nanowire field-effect transistor as a UV photodetector; optimization for maximum
sensitivity,” Phys. Status Solidi., A Appl. Mater. Sci. 206(1), 179–182 (2009).
E. Lee, D. I. Moon, J. H. Yang, K. S. Lim, and Y. K. Choi, “Transparent Zinc Oxide Gate Metal - Oxide Semiconductor Field-Effect Transistor for High-Responsivity Photodetector,” IEEE Electron Device Lett. 30(5),
493–495 (2009).
A. Zhang, C. Soci, B. Xiang, J. Park, D. Wang, and Y. H. Lo, “High Gain ZnO Nanowire Phototransistor,” in
Conference on Lasers and Electro-Optics (Baltimore 2007), 1–2.
G. Cheng, X. Wu, B. Liu, B. Li, X. Zhang, and Z. Du, “ZnO nanowire Schottky barrier ultraviolet photodetector
with high sensitivity and fast recovery speed,” Appl. Phys. Lett. 99(20), 203105 (2011).
Y. H. Ahn and J. Park, “Efficient visible light detection using individual germanium nanowire field effect
transistors,” Appl. Phys. Lett. 91(16), 162102 (2007).
K.-S. Shin, A. Pan, and C. O. Chui, “Channel length dependent sensitivity of Schottky contacted silicon
nanowire field-effect transistor sensors,” Appl. Phys. Lett. 100(12), 123504 (2012).
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12573
15. H. Yamamoto, K. Taniguchi, and C. Hamaguchi, “High-Sensitivity SOI MOS Photodetector with SelfAmplification,” Jpn. J. Appl. Phys. 35(Part 1, No. 2B), 1382–1386 (1996).
16. Y. Nakamura, H. Ohzu, M. Miyawaki, N. Tanaka, and T. Ohmi, “Design of Bipolar Imaging Device (BASIS),”
IEEE Trans. Electron. Dev. 38(5), 1028–1036 (1991).
17. Atlas, Silvaco., www.silvaco.com
18. R. F. Pierret, Semiconductor Device Fundamentals, (Pearson Education, 1996), Chap. 10, 11.
19. A. Fadavi-Roudsari, S. S. Saini, N. O, and M. P. Anantram, “High-Gain, Multiple-Gate Photodetector with
Nanowires in the Channel,” IEEE Electron Device Lett. 32(3), 357–359 (2011).
20. The results are obtained by solving the classic Poisson’s and carrier continuity equations using Atlas simulator
[17]. Room temperature dark currents are obtained by running simulations at higher temperatures, and then
estimating the current at room temperature. Electron and hole lifetimes are assumed to be 10-5s
21. M. Otto, M. Kroll, T. Käsebier, R. Salzer, A. Tünnermann, and R. B. Wehrspohn, “Extremely low surface
recombination velocities in black silicon passivated by atomic layer deposition,” Appl. Phys. Lett. 100(19),
191603 (2012).
22. O. Demichel, V. Calvo, A. Besson, P. Noé, B. Salem, N. Pauc, F. Oehler, P. Gentile, and N. Magnea, “Surface
recombination velocity measurements of efficiently passivated gold-catalyzed silicon nanowires by a new optical
method,” Nano Lett. 10(7), 2323–2329 (2010).
23. R. Coustel, Q. Benoît à la Guillaume, V. Calvo, O. Renault, L. Dubois, F. Duclairoir, and N. Pauc,
“Measurement of the Surface Recombination Velocity in Organically Functionalized Silicon Nanostructures:
The Case of Silicon on Insulator,” J. Phys. Chem. C 115(45), 22265–22270 (2011).
1. Introduction
Nanowires, with their miniaturized one dimensional size offer a variety of properties that are
of interest in different applications, including photodetectors. Quantum mechanical effects,
such as indirect to direct bandgap transition, prove useful in increasing the absorption
coefficient in nanowires [1]. The large surface to volume ratio of nanowires leads to a strong
capacitive coupling in metal-oxide-semiconductor geometries; and offers a better control over
the charge flow within the nano-scale channel of such structures. One can also take advantage
of the surface states in increasing the minority carrier recombination time [2].
The small size of nanowires limits their ability to absorb light efficiently [3].Therefore,
individual nanowires are preferred to be incorporated in photodetector geometries with
optical gain, such as phototransistors. The transistor can be a conventional junction bipolar or
a Metal Oxide Semiconductor (MOS) type [4, 5], or even a junction-less geometry [6].
A Junction-less transistor is formed by adding a gate to control a nanowire channel that
connects the source to the drain. The gate is able to create a potential barrier within the
nanowire channel, by depleting the nanowire out of the majority carriers. The height of the
potential barrier controls the amount of charge flow and the device ON-OFF state [6]. Such
structure offers the advantage of simpler fabrication, as: (i) all the extra steps required for
doping the source and the drain regions in a top-down approach are eliminated. (ii) Using
Silicon on Insulator (SOI) substrate and electron beam lithography/ etching steps allows easy
fabrication and localization of the nanowires as opposed to bottom-up approaches; and (iii)
SOI substrates facilitate the creation of the gate terminal by providing the buried oxide as the
gate dielectric and the bottom silicon substrate as the gate [7–9], although it is possible to
pattern and create a third terminal on top of the channel [10].
Junction-less transistor geometries are used for photo-detection applications, as well.
Examples are n- and p-type Si nanowire based photodetectors [8], Si, and ZnO nanowire
detectors with surface defects that separate the carriers and increase the recombination
lifetime [2, 11, 12], and Ge, ZnO, Si, and GaN nanowire based photodetectors with back
gates [7, 9, 13, 14]. As a photodetector, the junction-less transistor can be biased in either ON
or OFF state. However, the focus of this manuscript is the OFF state, as the dark current and
its associated shot noise are both low. This gives the opportunity of detecting low level
intensities. The operating principle of such a phototransistor is similar to the lateral bipolar
action [15]. When the channel is illuminated, most of the photo-generated minority carriers
easily travel towards the contacts in response to the high electric field. However, the majority
carriers are trapped inside the potential barrier created by the gate (OFF state). To satisfy the
charge neutrality the barrier height is reduced. A lower potential barrier creates a pathway for
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12574
the trapped carriers to diffuse; but more importantly it eases the overall carrier flow and
pushes the transistor towards ON state. The overall photocurrent is therefore increased.
A low level of dark current is a great advantage of junction-less phototransistors over
photoconductors. However, lack of junctions can generally lead to a low optical gain. Due to
the small cross sectional area of nanowires, the absorption in nanowires and therefore the
quantum efficiency of the device would also be low [3]. Both of these issues could lead to a
poor noise equivalent power (NEP) for the phototransistor. In this study, we propose ways to
improve the performance of the junction-less, silicon based phototransistor in terms of optical
gain and sensitivity, by modifying the channel geometry and band-structure engineering by
using multiple gates. Such a structure is able to potentially detect very low light intensities.
2. Phototransistor design: role of nanowires
Figure 1(a) illustrates the proposed structure. The channel in this structure (top view) is partly
wide for light absorption and partly narrow to lower the dark current (NW1 that is covered by
the ‘primary’ gate). We also modify the source region by adding a second nanowire, NW2,
and a ‘secondary’ gate that allows for controlling the carrier concentration in the source,
without the need for changing the doping concentration in the source region, as will be
discussed shortly.
For discussion and demonstration purposes, we use the wavelength of 630nm in this work.
The semiconductor in this study is p-type, and therefore the primary gate has to be biased
positively to keep the phototransistor in OFF state. The total channel length is 5µm. The
channel is 2µm wide everywhere except at the nanowire regions. The gate oxide thickness is
20nm. Such photodetector can be integrated into a pixel by being connected to a high
impedance integration node. A capacitor at this node converts the photocurrent into a voltage
value that is read out later [4, 16].
The first key element in our design is the channel. In junction-less transistors, the channel
is normally very thin to ensure the gate is able to fully deplete it. Considering the relatively
small absorption coefficient of silicon at the wavelength of interest (≈3.9x103cm−1 [17]), the
drawback of the thin absorption region is low external quantum efficiency. Therefore, we try
to increase the thickness of the semiconductor layer, while making sure the gate is still strong
enough to fully deplete the channel. This is achieved by replacing the channel underneath the
primary gate with a narrow layer, as marked with NW1 in Fig. 1(a). Use of NW1 allows for
introducing the gate in three dimensions, which helps to complete the depletion of the
channel. Furthermore, since the gate control over NW1 is strong, the channel and the gate can
be short, while still avoiding short channel effects. Narrow, short regions can also lead to a
smaller capacitance imposed by the gate, and speed up the device operation compared to long
gates.
The rest of the channel on the drain side is kept wide, to provide a surface for absorption
of light. Furthermore, the position of the primary gate is towards areas closer to the source, to
increase the surface for photon absorption between gate and drain, similar to the base and
collector junction in a bipolar phototransistor [18]. We chose a silicon thickness of 0.85µm
for this work that would result in a quantum efficiency of about 20%, assuming the light that
is illuminated on the top of the device, is mainly absorbed over the wide areas. The surface
reflection is also assumed to be about 0.3. A thicker semiconductor layer would of course
lead to higher quantum efficiency. However, considering the nanowire widths of a few
hundred nanometers and narrower, fabricating the gate contact over such high aspect ratio
channel can be challenging.
The next key modification comes from the concept of bipolar transistors. In a welldesigned bipolar transistor, to push the emitter efficiency towards unity, the doping level of
the emitter (source) has to be much larger than the base (channel) [18]. To satisfy this
condition, instead of actually doping the source region with acceptors, we use a ‘secondary’
gate to control the carrier concentration of the source, the same way we use the ‘primary’ gate
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12575
to control the carrier concentration over the channel. The secondary gate is located in an area
between the primary gate and the source contact. For a p-type semiconductor the secondary
gate is biased negatively in order to increase the concentration level of holes over the source
region. Since the impact of the gate in carrier accumulation is local, in order to make the gate
effect more tangible, the area underneath the secondary gate is narrowed down, as shown by
NW2 in Fig. 1(a). We note that the narrow regions of NW1 and NW2 are more like narrow
slabs and not nanowires in thick devices. However, the ultimate goal is to replace them with
nanowires that bridge the thick, wide areas of channel and source.
Drain
A
NW1
NW2
Source
Primary gate
A’
Secondary gate
Source
Primary gate
Secondary gate
Drain
Drain
Buried oxide
(a)
1
18
Primary gate
-3
Carrier concentration (cm )
16
Energy (eV)
0.5
0
Secondary gate
−0.5
−1
0
Conduction band
Valance band
Fermi level
1
2
14
12
10
8
6
4
Primary gate
X (μm)
(b)
3
4
5
2
0
Holes
Electrons
1
Secondary gate
2
X (μm)
3
4
5
(c)
Fig. 1. (a) Junction-less phototransistor with multiple gates. The semiconductor is entirely ptype (doping: 1015cm−3), with the thickness of 0.85µm. The total channel length is 5µm. The
channel is 2µm wide everywhere except the narrow regions. The gate oxide is 20nm (not
shown in figure). The secondary gate is lifted up to show the narrow region NW2. A similar
geometry is present underneath the primary gate that is shown in the top view, where the gates’
top layer is removed. Light is shined through a window with the area of 5.6µm2. (b), (c)
Energy band diagram and carrier concentration (logarithmic scale) of the structure, along
cutline AA’, located about 100nm below the top silicon and oxide interface. Dashed lines
represent the energy band and carrier concentration when the secondary gate is removed.
VGprimary = 1V, VGsecondary = −1V, VS = 0V, VD = 0V.
The conduction and valance bands of the device, together with the Fermi level are plotted
in Fig. 1(b). The data is obtained along the channel, shown by cutline AA’ in the top view of
Fig. 1(a). To obtain this data, the primary gate is biased at +1V, while the secondary gate is
biased at −1V. The source and drain contacts are both connected to the ground. The energy
band diagram clearly shows the resemblance of the device to a pnp bipolar transistor. The
electron and hole concentration of the structure in Fig. 1(c) also confirm that electrons are the
majority carriers in the channel area under the primary gate; while holes are the majority
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12576
carriers in the source and drain regions. The secondary gate increases the concentration of the
majority carriers in the source region, similar to the high doping level at the emitter of a
junction bipolar transistor. If the secondary gate is removed, the barrier height on the sourcechannel region will decrease, as marked with dashed lines in Fig. 1(b). Similarly the
concentration of holes on the source side declines, as plotted in Fig. 1(c).
Figures 1(b) and 1(c) clearly demonstrate the importance of both gate/nanowire
geometries in our design. The primary gate/ NW1 combination is responsible for converting a
simple photoconductor into a phototransistor; while the role of the secondary gate/NW2 is to
control the optical gain. Since other works, such as [9] have illustrated the role of gate/
nanowire geometry (primary gate/ NW1 in our design) on the output parameters, the focus of
this work is investigating the impact of NW2 on the performance parameters of the
phototransistor.
We note that there are significant differences between the device proposed here and the
one in reference [19]. The lateral bipolar action in each device is maintained differently, due
to the fact that the operating principles of the junction and junction-less transistors are
different. In [19], the OFF state was obtained via the reverse biased drain-channel p-n
junction and the primary gate that accumulated the carriers within the channel. Whereas here,
the potential barrier during the OFF state is created only by the primary gate that pushes the
channel towards depletion and inversion. In both structures, the nanowire/secondary gate
combination is used to increase the optical gain; but the implementation is different. The
secondary gate in [19] was added over the channel for the purpose of lowering the barrier by
depleting the carriers. Here the role of the secondary gate is to artificially create a junction.
This eliminates the need for having a doped p-n junction, and simplifies the design of the
junction-less phototransistor.
3. Results and discussion
Figure 2(a) shows how the source current is influenced by the voltage of the secondary gate
[20]. Figure 2(b) shows the conduction band energy of the structure, along the channel, for
different voltages of the secondary gate. The primary gate is biased at +1V to create the OFF
state potential barrier. To emphasize the importance of the secondary gate, we have plotted
the current at both positive and negative values of secondary gate. When the secondary gate is
biased positively, it acts like an extension of the primary gate. The device exhibits a low level
of dark current, as it is biased in the OFF state. The current is increased when the device is
illuminated. However, in comparison with the case of dark, the barrier height of the
illuminated structure shows a very slight change (Fig. 2(b)), suggesting that the optical gain is
small, and in fact calculations show an optical gain of less than 1.
As the secondary gate bias is shifted towards more negative values, the drain current starts
to increase. The conduction band energy in Fig. 2(b) shows an increase in the carrier
concentration at the source region, as if it is doped with a higher concentration of acceptors.
This directly impacts the emitter efficiency of the structure, that is rephrased below, based on
the definition of the emitter efficiency in a one dimensional, pnp junction transistor [18]:
γ=
Ip
I p + In
(1)
Reinterpreting the definition of the emitter efficiency to suit the junction-less transistor
case, Ip denotes the current of the holes, diffusing from source into the channel; and In is the
current of the electrons that are injected into the source from the channel. When the
concentration of holes in the source is increased, the holes that diffuse into the channel
outnumber the channel electrons that diffuse into the source. The ratio of Ip/In would therefore
increase, and the emitter efficiency approaches the limit of 1. As a result, one expects a larger
current.
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12577
10
-8
Dark
Conduction band energy (eV)
IS (A)
10
-10
10
-12
10
-14
10
-16
VG
=.5V,
VG2nd
=.5V, dark
dark
2nd
1.2
0.1mW/cm2
VG
=.5V,
VG2nd
=.5V, illumination
illumination
2nd
1
VG
VG2nd
==−1V,
1V, dark
dark
2nd
VG2nd==−1V,
illumination
VG
1V, illumination
0.8
2nd
Secondary gate
0.6
0.4
0.2
A
0
A’
- 0.2
- 0.4
-2
- 1.5
-1
- 0.5
VGsecondary (V)
(a)
0
0.5
1
0
Primary gate
1
2
X (μm)
3
4
5
(b)
Fig. 2. (a) Source current versus the secondary gate bias in the junction-less phototransistor.
The primary gate is 200nm long and covers a 200nm wide channel. The secondary gate is
1.2um long, and covers a 20nm wide region (NW2). (b) Impact of the secondary gate bias over
conduction band energy of the structure along the channel (cutline AA’). VGprimary = 1.0V, VS
= 0.5V, VD = 0V.
Although the secondary gate bias affects both dark, and photo- currents, the change in the
current is less pronounced at dark. As illustrated in Fig. 2(a), at dark the current changes by
about 40 times when the secondary gate is changed from +1V to −2V. For the same
conditions, the photocurrent changes more than 700 times. This is due to the fact that at dark
the device is in its OFF state, even when the secondary gate is negatively biased. Under such
condition, the potential barrier is slightly higher than the case when the secondary gate is
positive (Fig. 2(b)). This change in the potential barrier is responsible for the larger dark
current at negative bias of the secondary gate. However, the source-channel junction is still
reverse biased and the current change is low as a result. Whereas under illumination and when
the secondary gate is negative, the source-channel junction is forward biased, allowing for a
larger current to pass. The potential barrier of the structure shows a more dramatic change
under illumination, confirming that the detector has a larger emitter efficiency. As a result,
the dark current increase is smaller than the increase in photo current, in agreement with the
numerical results in Fig. 2(a).
Narrowing the size of NW2 can similarly lead to an increase in the current. This is verified
in Fig. 3(a), where the source current versus the width of NW2 is plotted. The current change
is attributed to fact that accumulation of holes under the secondary gate is a local effect. We
have plotted the conduction band energy for two source widths of 100nm, and 20nm in Fig.
3(b). The energy band is obtained along a cutline across the source. When the source is wide,
only the areas that are very close to the interface are in accumulation. As the source width is
narrowed down, a larger percentage of the channel is influenced by the secondary gate, and
the net concentration of holes is increased. The emitter efficiency, and therefore the source
current will both increase.
Similar to the case of the secondary gate bias, the influence of the width of NW2 on the
source current is less at dark condition. For example, in Fig. 3(a) the source current changes
by more than two orders of magnitude when the device is illuminated; while the dark current
varies by about 25 times. This behavior is also due to the smaller emitter efficiency of the
device at dark, as discussed previously.
Investigation of the role of the length of NW2 is plotted in Fig. 4(a). We note that the total
length of the detector is kept fixed. Elongating NW2 causes the source current to increase;
however, it does not change the barrier height under the secondary gate (Fig. 4(b)). The
change in the current is attributed to the funneling effect [19]. The long, narrow diffusion
pathway creates a physical barrier for electrons and slows down the diffusion process. This
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12578
forces the potential barrier under the primary gate to decrease; which in turn causes the
photocurrent to increase.
10
-10
0.53
Dark
B
2
-12
IS (A)
10
Conduction band energy (eV)
0.1mW/cm
10
10
-14
-16
0
0.2
0.4
0.6
Width of NW2 (μm)
0.8
1
W=20 nm
W=100 nm
B’
0.52
0.51
0.5
0.49
0.48
0.47
0.46
0.94
0.96
(a)
0.98
1
X (μm)
1.02
1.04
1.06
(b)
Fig. 3. (a) Source current versus nanowire width underneath the secondary gate. The primary
gate is 200nm long and covers a 200nm wide nanowire. The secondary gate is 400nm long.
VGprimary = 1.0V, VS = 0.5V, VD = 0V, VGsecondary = −1.0V. (b) Conduction band energy,
across the source area (cutline BB’).
10
-9
10
-10
10
-11
10
-12
10
-13
10
-14
10
-15
0.9
L=0.2 μm
L=1.2 μm
0.8
Conduction band energy (eV)
IS (A)
By combining the three parameters discussed in this section, the secondary gate bias, the
width, and the length of NW2, it is possible to improve the performance of the junction-less
phototransistor in terms of illumination current and noise equivalent power. We have
tabulated a few cases in Table 1. The data ranges from different biases of the secondary gate,
to different widths and lengths of NW2. The most dramatic increase in the optical gain takes
place where the width of the nanowire is reduced from 400nm (first row) to 20nm (second
row). This verifies the importance of having a narrow region under the secondary gate.
Decreasing the voltage and increasing the length of the secondary in rows 3 and 4 are both
effective, but the improvement is not as much as the change in the width of NW2. We note
that for the data presented in the table the secondary gate is biased in accumulation, which is
the best possible bias as verified in Fig. 2(a).
Dark
0.1mW/cm2
0.2
0.4
0.6
0.8
1
Length of secondary gate (μm)
(a)
1.2
0.7
0.6
0.5
0.4
0.3
A
A’
0.2
0.1
0
0
1
2
X (μm)
3
4
5
(b)
Fig. 4. (a) Source current versus nanowire length underneath the secondary gate. (b)
Conduction band energy, along the channel for different lengths of the secondary gate and
NW2. For both cases the primary gate is 200nm long and covers a 200nm wide nanowire, and
the secondary gate covers a 20nm wide channel. VGprimary = 1.0V, VGsecondary = −1.0V, VS =
0.5V, VD = 0V.
We have also included a case in which the thickness of NW2 is reduced (fifth row), as if a
nanowire bridges the channel and the source regions. The effect of decreasing the thickness of
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12579
NW2 is very similar to the effect of decreasing its width. It strengthens the control of the
secondary gate over NW2 and increases the concentration of majority carriers. As a result, one
expects a better emitter efficiency and photocurrent. Although the case of the narrow, thin
NW2 seems to be challenging in terms of fabrication, it exhibits the best photocurrent and
NEP among the other devices in the table.
Table 1. Comparison of Multiple Gate Phototransistors
Device specifications1
IDark (A)
IPhoto2(A)
Optical
gain
Responsivity
(A/W)
NEP3
(W/Hz0.5)
Photon
flux
(cm-1)
Width (nm)
400
Thickness (µm)
0.85
2.7×10-16 4.7×10-13
0.8
0.08
1.8×10-16
588
Length (µm)
0.4
G2nd
Bias (V)
-1
Width (nm)
20
NW2
Thickness (µm)
0.85
2
5.9×10-15 6.6×10-11
114.8
11.7
5.6×10-17
180
Length (µm)
0.4
G2nd
Bias (V)
-1
20
Width (nm)
NW2
Thickness (µm)
0.85
3
264.3
27.0
3.7×10-17
118
5.9×10-15 1.5×10-10
1.2
Length (µm)
G2nd
Bias (V)
-1
Width (nm)
20
NW2
Thickness (µm)
0.85
4
7.2×10-15 2.9×10-10
512.7
52.4
2.9×10-17
94
Length (µm)
1.2
G2nd
Bias (V)
-2
20
Width (nm)
NW2
Thickness (µm)
0.05
4.0×10-15 1.2×10-09
5
2053.1
209.7
1.1×10-17
35
1.2
Length (µm)
G2nd
Bias (V)
-2
1
Width of NW1=200nm, Length of primary gate=200nm, Semiconductor thickness=0.85µm, VGprimary=1.0V
2
A surface reflection of 30% is considered. The photocurrent presented here is multiplied to transmission.
3
The noise sources considered in calculating the NEP are input and source shot noise, and also the thermal noise of
the channel. We did not calculate the Flicker noise, as it strongly depends on how the device is fabricated. The load
is considered to be infinitely large; otherwise the thermal noise of the load at room temperature dominates.
NW2
1
We remark that for the results presented in Table 1 the effect of the surface states is not
included, while the large surface to volume ratio in nanowires can lead to carrier
recombination, and therefore degradation of the photodetector response. To study this, we
chose the devices on rows 2 and 5 of Table 1, and investigated their response while the effect
of surface states included. The surface recombination velocity of passivated silicon nanowires
ranges from less than 13cm/s to about 61cm/s in the state of the art structures reported in
literature [21–23]. In our simulations, the surface recombination velocity is assumed to be
13cm/s. As a result of surface recombination, the photocurrent of the device on the second
row of Table1 has dropped to 3.6 × 10−11A. The device in the fifth row of the table has a
thinner NW2, which translates into a higher surface to volume ratio. The photocurrent of this
device is decreased to 1.45 × 10−10A. The decrease in the photocurrent is due to the
recombination of carriers within both NW1 and NW2 regions. In NW1, carrier recombination
leads to losing some of the photo-generated carriers. As a result, the barrier would not
decrease as much as in the ideal case, leading to degradation in the overall current flow.
Recombination of carriers as they pass through NW2 is also responsible for further decrease
in the current. Work to decrease the surface recombination velocity and bulk recombination
times will be helpful in increasing the optical gain but as expected they will lead to an
increase in the time scales associated with the transient response.
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12580
The speed of the photodetector depends on a different set of physical processes during the
rise and fall of photocurrent. When the light source is turned ON, the rise time (speed) is
limited by the RC time constant. The capacitance (C) is determined by the two gates, and the
resistance (R) is determined by the narrow NW regions. Since in the devices presented, NW2
is longer and narrower than NW1, we expect the time constant of this region to dominate.
Once the light source is switched OFF, the extra carriers in the NW1 region recombine and
increase the potential barrier (under the primary gate) to its original value. In this case the size
of NW2 does not cause much variation in the fall time, as the dominant process is the carrier
recombination in NW1.
The transient response of a structure with NW1 = 200nm, and NW2 = 20nm is plotted in
Fig. 5. The logarithmic scale in the inset shows that when the light is switched OFF, the
current drops by about 3 orders of magnitude in the first few milliseconds. Normally external
circuits in a pixel apply a reset signal to the photodetector to decrease the fall time. We have
also summarized the rise time and fall time of two more structures in Table 2. The data shows
a small difference in values as the size of NW2 is changed.
x 10
1
-10
0.9
0.8
0.7
Light ON
Light OFF
Light ON
10
-9
10
-10
10
-11
10
-12
0.3
10
-13
0.2
10
-14
0.1
10
-15
IS (A)
0.6
0.5
IS (A)
0.4
0
0.5
1
1.5
Time, t (s)
0
0
0.5
1
1.5
Time, t (s)
Fig. 5. Transient response of a junction-less, multiple gate photodetector, with NW1 = 200nm,
and NW2 = 20nm. The inset shows the same data in logarithmic scale. VGprimary = 1.0V,
VGsecondary = −1.0V, VS = 0.5V, VD = 0V, Intensity = 10−4W/cm2.
Table 2. Rise Time and Fall Time of Multiple Gate Phototransistors
Width/ Length/ thickness
of NW2 (nm/µm/µm)
Rise time (ms)
Fall time (ms)
1000/0.4/0.85
20/0.4/0.85
20/1.2/0.05
0.8
0.7
1
1
1.1
0.95
Notes:
1
Rise time: the time the current chnages from 10% to 90% of the peak value.
2
Fall time: the time the current drops from 90% to 10% of the peak value.
3
Width of NW1 = 200nm, Length of primary gate = 200nm.
4
VGprimary = 1.0V, VS = 0.5V, VD = 0V, Intensity = 10−4W/cm2.
5
First two rows: VGsecondary = −1.0V; third row VGsecondary = −2.0V.
4. Conclusion
In summary, we proposed a new geometry for junction-less phototransistors. In this design,
instead of having an entirely narrow channel, we keep the channel partly narrow and partly
wide for electrostatic and optical reasons, respectively. The narrow region is covered by a
primary gate that creates the potential barrier required for phototransistor operation. The wide
area of the channel provides a surface for better light absorption. Further boost of the optical
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12581
gain is achieved by increasing the emitter efficiency of the device, by adding a secondary gate
and narrowing down the source region, which artificially increases the majority carrier
concentration in the source. This in turn causes the optical gain to increase and the NEP to
improve. We demonstrated multiple gate photodetectors showing NEP of 1.1x10−17W/Hz0.5,
which corresponds to a photon flux of about 35s−1.
Acknowledgments
The authors acknowledge support from National Science Foundation (NSF) grant 1001174,
and Waterloo Institute for Nanotechnology (WIN) fellowship. We would like to thank M.
Keshavarz Akhlaghi and D. Shiri for their valuable suggestions.
#207210 - $15.00 USD
(C) 2014 OSA
Received 25 Feb 2014; revised 4 May 2014; accepted 5 May 2014; published 16 May 2014
19 May 2014 | Vol. 22, No. 10 | DOI:10.1364/OE.22.012573 | OPTICS EXPRESS 12582
Download