ArticlePDF Available

Large-area low-jitter silicon single photon avalanche diodes

Authors:

Abstract and Figures

Single photon counting (SPC) and time correlated single photon counting (TCSPC) techniques have been developed in the past four decades relying on photomultiplier tubes (PMT), but interesting alternatives are nowadays provided by solid-state single photon detectors. In particular, silicon Single Photon Avalanche Diodes (SPAD) fabricated in planar technology join the typical advantages of microelectronic devices (small size, ruggedness, low operating voltage and low power dissipation, etc.) with remarkable basic performance, such as high photon detection efficiency over a broad spectral range up to 1 mum wavelength, low dark count rate and photon timing jitter of a few tens of picoseconds. In recent years detector modules employing planar SPAD devices with diameter up to 50 µm have become commercially available. SPADs with larger active areas would greatly simplify the design of optical coupling systems, thus making these devices more competitive in a broader range of applications. By exploiting an improved SPAD technology, we have fabricated planar devices with diameter of 200 mum having low dark count rate (1500 c/s typical @ -25 °C). A photon timing jitter of 35 ps FWHM is obtained at room temperature by using a special pulse pick-up network for processing the avalanche current. The state-of-the-art of large-area SPADs will be reviewed and prospects of further progress will be discussed pointing out the challenging issues that must be faced in the design and technology of SPAD devices and associated quenching and timing circuits.
Content may be subject to copyright.
Large-area low-jitter silicon single photon avalanche diodes
Massimo Ghioni
*
a, b
, Angelo Gulinatti
a
, Ivan Rech
a
, Piera Maccagnani
c
, and Sergio Cova
a, b
a
Politecnico di Milano, Dipartimento di Elettronica e Informazione, Piazza Leonardo da Vinci 32 -
20133 Milano, Italy
b
MPD Micro-Photon-Devices, via Stradivari 4 – 39100 Bolzano, Italy
c
IMM-CNR sezione di Bologna, Via Piero Gobetti, 101 – 40129 Bologna, Italy
ABSTRACT
Single photon counting (SPC) and time correlated single photon counting (TCSPC) techniques have been developed in
the past four decades relying on photomultiplier tubes (PMT), but interesting alternatives are nowadays provided by
solid-state single photon detectors. In particular, silicon Single Photon Avalanche Diodes (SPAD) fabricated in planar
technology join the typical advantages of microelectronic devices (small size, ruggedness, low operating voltage and low
power dissipation, etc.) with remarkable basic performance, such as high photon detection efficiency over a broad
spectral range up to 1 µm wavelength, low dark count rate and photon timing jitter of a few tens of picoseconds. In
recent years detector modules employing planar SPAD devices with diameter up to 50 µm have become commercially
available. SPADs with larger active areas would greatly simplify the design of optical coupling systems, thus making
these devices more competitive in a broader range of applications. By exploiting an improved SPAD technology, we
have fabricated planar devices with diameter of 200 µm having low dark count rate (1500 c/s typical @ -25 °C). A
photon timing jitter of 35 ps FWHM is obtained at room temperature by using a special pulse pick-up network for
processing the avalanche current. The state-of-the-art of large-area SPADs will be reviewed and prospects of further
progress will be discussed pointing out the challenging issues that must be faced in the design and technology of SPAD
devices and associated quenching and timing circuits.
Keywords: single photon avalanche diodes, SPAD, time-correlated single photon counting, TCSPC, photon timing
1. INTRODUCTION
There is nowadays a widespread and steadily growing interest in single photon detectors, driven by the need for ultimate
sensitivity in various scientific and industrial applications such as fluorescence spectroscopy in life and material
sciences, quantum cryptography and computing, profilometry of remote objects with optical radar techniques, particle
sizing etc. In particular, the use of fluorescence lifetime spectroscopy as both an analytical and research tool has
increased markedly in recent years.
To date fluorescence lifetime spectroscopy has found remarkable applications in
chemistry, biochemistry and
biology (for a review, see for example [1]). Most biologically relevant fluorophores exhibit
characteristic decay times ranging from picoseconds to nanoseconds. The
time-correlated single-photon counting (TCSPC)
technique [2, 3] is currently used for directly measuring fluorescence decays.
TCSPC was developed relying on
photomultiplier tubes (PMTs), that is, vacuum tube detectors with high internal gain. High performance PMTs have been
produced industrially with sophisticated technologies since the 60’s; amongst their advantages, the most valuable one is
the wide sensitive area ( cm
2
), which in some cases greatly simplifies the design of the optical system. PMTs also attain
remarkable performance at high counting rate and offer picosecond timing resolution with micro channel plate (MCP)
models. However, they suffer from low quantum efficiencies (QE) in the visible range: the QE of conventional bialkali
and multialkali photocathodes reaches 20-25% between 400 and 500 nm [4].
*
ghioni@elet.polimi.it, phone +39-02-23996093, fax: +39-02-2367604
Invited Paper
Quantum Sensing and Nanophotonic Devices V, edited by Rengarajan Sudharsanan, Christopher Jelen,
Proc. of SPIE Vol. 6900, 69001D, (2008) · 0277-786X/08/$18 · doi: 10.1117/12.761578
Proc. of SPIE Vol. 6900 69001D-1
2008 SPIE Digital Library -- Subscriber Archive Copy
In more recent years, single photon avalanche diodes (SPADs) have emerged as a solid state alternative to PMTs [5-8].
Besides the well known advantages of solid state versus vacuum tube devices (small size, ruggedness, low power
dissipation, low supply voltage, high reliability, etc.), SPADs provide inherently higher quantum efficiency, particularly
in the red and near infrared spectral regions. The small active area of SPAD devices is often considered a disadvantage
compared to PMTs. However, this is not true in a variety of TCSPC techniques relying on descanned confocal or near
field microscopy such as fluorescence lifetime imaging (FLIM) and time-resolved fluorescence correlation spectroscopy
[1, 3]. In these applications, the spot size at the microscope image plane can easily be made underfill the SPAD active
area, provided that the diameter is sufficiently large ( 50 µm). In confocal microscopy, the SPAD device can also be
used as confocal pinhole, thus simplifying the whole optical system and increasing the miniaturization potential.
Although active area diameters between 50 and 100 µm are acceptable in the aforementioned applications, larger
diameters (> 100 µm) would be highly desirable for attaining good photon collection efficiency without requiring
complex and time-consuming optical alignment and focusing procedures. Furthermore, fiber pigtailing of the detector,
often employed for having a more flexible optical system, would be simpler and more efficient for SPAD detectors with
large area.
Unfortunately, large active area, low dark counting rate, high quantum detection efficiency and small photon timing jitter
are conflicting requirements for a SPAD device [9].
Single photon counting modules based on SPAD diodes with thick depleted region (20-25 µm) and large active area (180
µm diameter) are commercially available from PerkinElmer Optoelectronics [10]: they have very good photon detection
efficiency (PDE) and low dark counting rate (DCR), but they have a moderate photon timing resolution of typically 350
ps full-width at half maximum (FWHM). Such a performance is unsuitable for tissue autofluorescence measurements
[11] or protein conformational dynamic analysis [12], where fluorescence components as short as 100 ps or less are often
met.
SPAD devices with thin depleted region (1-2 µm), fabricated in a standard CMOS silicon technology can attain time
resolution down to less than 50 ps FWHM at room temperature [13, 14]. This performance is comparable with that of the
best MCP-PMTs. Standard CMOS technology makes it possible to monolithically integrate the SPAD device and its
associate quenching circuit, taking great advantage of reduced size and low parasitic capacitances. Unfortunately, CMOS
technologies available in industrial silicon foundries are not optimized for SPAD device fabrication, resulting in reduced
photon detection efficiency and relatively high dark counting rate. So far, no CMOS-based SPAD devices with active
area diameter > 100 µm have been reported.
To overcome these limitations and provide a solid-state alternative to MCP-PMTs in demanding TCSPC applications we
have developed an improved planar silicon process that enables the fabrication of high-performance SPAD devices with
active area diameter up to 200 µm.
2. DEVICE TECHNOLOGY ISSUES
It is well known that semiconductor detectors offer better photon detection efficiency (PDE) than vacuum tube detectors
since electron emission phenomenon is not involved and the primary photocurrent flows directly within the
semiconductor. However, the situation is just opposite from the standpoint of the detector noise, that is, of the detector
dark current or dark count rate (DCR). Vacuum tube detectors have dark current per unit of sensitive area inherently
lower than semiconductor detectors. For instance, S20 photocathodes are currently obtained with thermal electron
emission lower than 1000 electrons s
-1
cm
-2
at room temperature. On the other hand, the dark current in a reverse biased
silicon junction operating at room temperature or below is dominated by thermal generation of carriers in the depletion
layer and tunneling processes [15]. The total generation rate G per unit volume is given by:
G = G
th
+ G
TAT
+ G
bbt
(1)
where G
th
is the contribution due to thermal generation via localized deep energy levels (also called traps), G
TAT
is the
trap-assisted tunneling contribution and G
bbt
is the band-to-band tunneling contribution [15-17]. It has been shown
theoretically and experimentally demonstrated that G
bbt
becomes the dominant contribution only for electric field
strengths higher than 710
5
V/cm [16, 17]. We will show in the next section that this contribution can be made negligible
by properly reducing the peak electric field within the active area of the SPAD device. In this condition, the total
generation rate per unit volume can be written as:
pn
pn
TTATth
ee
ee
NGGG
+
=+
(2)
Proc. of SPIE Vol. 6900 69001D-2
E - Et
E
Pure
tunneling
where N
T
is the volume density of deep levels, e
n
and e
p
are the probabilities per unit of time for the emission of an
electron or a hole respectively. Both the quality of the starting material and the technological processes used in the
device fabrication have a strong impact on N
T
and therefore on the generation rate G. Transition metal impurities are the
most common source of deep levels. Metal contamination may occur during silicon handling, high-temperature heat
treatments or ion implantations As unintentional contaminants Fe, Cu, Ti or Ni are usually found in silicon in
concentrations of 10
11
-
10
12
cm
3
. On the other hand, the generation rate is strongly dependent on the electric field
acting on deep levels as well [16-19]. Three different mechanisms cause the increase of emission rates e
n
and e
p
from
deep levels under the applications of strong electric fields: the Poole-Frenkel effect, the phonon-assisted tunneling, and
the direct tunneling from deep level into conduction or valence bands (Fig. 1a, b).
Fig.1 Potential barrier for the emission of an electron from a deep energy level in external electric field for: a) charged impurities
and b) neutral impurities. The arrows show the field-enhanced emission processes.
Fig. 2 Schematic diagram showing the field-enhanced emission processes of electrons and holes from donor-like (a) and acceptor-
like (b) deep levels. Either electron emission from a donor-like level or a hole emission from an acceptor-like level is
enhanced by the Poole-Frenkel effect.
The Poole-Frenkel (PF) effect consists of the lowering of the potential barrier due to the electric field applied to a
semiconductor. This effect is present only when the deep impurity level behaves like a Coulombic well, i.e., the emission
of electrons or holes ionizes the impurity. Phonon-assisted tunneling and direct tunneling are possible for impurities in
all charge states. Hence, carrier emission rate can increase in an electric field due to carrier tunneling even in the case of
a neutral impurity (Dirac well). Whether a deep level behaves like a Coulombic or a Dirac well during the emission of a
carrier it depends on its charge state (Fig. 2a, b). In practice, if the deep level is acceptor-like, i.e. neutral when empty, it
a)
b)
a) b)
Proc. of SPIE Vol. 6900 69001D-3
Electric field (V/cm)
0
04
10
ci,
E
10
102
101
106
E
V
a)
F
2
0
0
a)
E
a)
0
a)
V
a)
1000
Li
0.6 1.2
x(pm)
1.4 1.6
behaves like a Coulombic well during the hole emission and like a Dirac well during electron emission. Conversely, if
the deep level is donor-like, i.e. neutral when filled, it behaves like a Coulombic well during electron emission and like a
Dirac well during hole emission. To evaluate the impact of the electric field on the generation rate, we have considered a
donor-like and an acceptor-like deep level with energy E
T
=E
C
-0.28 eV, which are associated to Ti and FeB impurities in
silicon respectively [20]. Since these traps are closer to the conduction band, the limiting mechanism to the generation
rate is hole emission. Therefore:
pT
eNG
(3)
The field enhancement factor Γ for the generation rate is given by:
)0(e
)F(e
p
p
p
=Γ=Γ
(4)
where F is the electric field strength.
Fig. 3a shows the field enhancement factor Γ for the donor-like and for the acceptor-like deep levels as a function of the
electric field strength. The dependence of Γ on the position within the depletion layer of a typical SPAD junction is
shown in Fig. 3b, where the electric field profile is also shown for clarity. A remarkable Γ of 4000 is obtained for the
acceptor-like level with F ~ 6 10
5
V/cm, whereas the donor-like level shows a lower Γ of 70.
Fig. 3 Field enhancement factor Γ for a donor-like and an acceptor-like deep level with energy E
T
=E
C
-0.28 eV. a) Γ as a function
of the electric field strength. b) Γ as a function of the vertical position within the active region of a SPAD device.
Modern silicon technologies guarantee almost no structural defects and low metal contamination levels as well.
However, even trace levels of metal impurities that would be normally undetectable with state-of-the-art analytical
techniques might impair the DCR performance of SPAD detectors, due to the inherently high electric field within the
depletion layer.
Under the simplifying hypothesis of a single deep level impurity, the DCR expected from a SPAD can be calculated as
follows:
(
)
Aw
n
Aw
n
dx
)F()0(e)F()0(e
)0(e)0(e)F()F(
)0(e)0(e
)0(e)0(e
ANdx
)F(e)F(e
)F(e)F(e
ANDCR
eff,g
i
eff
0,g
i
x
x
T
ppnn
pnpn
pn
pn
T
x
x
T
pn
pn
T
p
n
p
n
τ
=Γ
τ
=
=
η
Γ+Γ
+ΓΓ
+
=
η
+
=
(5)
where
(
)
dx
)F()0(e)F()0(e
)0(e)0(e)F()F(
w
1
p
n
x
x
T
ppnn
pnpn
eff
η
Γ+Γ
+ΓΓ
=Γ
(6)
Proc. of SPIE Vol. 6900 69001D-4
n++ shallow n-i--i-
isolation
H
5pm
I-
p- epilayer
p -I- buried layer
4-substrate
hv
50-200pn,
—Th
Anode
Cathode
and
eff
0,g
eff,g
Γ
τ
=τ
(7)
Γ
eff
is an effective field enhancement factor, n
i
= 1.45 10
10
cm
-3
is the intrinsic carrier concentration at room temperature,
A is the junction area, w is the depletion region width,
τ
g.0
is the low-field generation lifetime (tens of ms in high quality
silicon wafers [21]),
η
T
is the avalanche triggering probability [22], x
n
and x
p
are the depletion edges. Equation 7 shows
that the effective generation lifetime in SPAD detectors is lower than the low-field generation lifetime by a factor
Γ
eff
where, in practical cases, the value of
Γ
eff
may range from tens to thousands depending on the electric field strength and
on the trap position and kind. Simple quantitative evaluations clearly show that:
a) for developing SPADs with low DCR and fairly wide area it is necessary to have a high quality fabrication
technology leading to a consistent reduction of impurity concentration with respect to standard, high-quality silicon.
In addition, the electric field profile within the depletion region must be properly designed for ensuring a low value
of
Γ
eff
;
b) better SPAD design and fabrication technology enable a significant increase of the active area of SPAD detectors.
SPAD active area will be anyway much smaller than that of PMTs having comparable level of noise.
A couple of examples about the fabrication of planar SPAD with depletion layer width w = 1 µm can well illustrate these
points. If we target a DCR not exceeding 1000 c/s in a detector with diameter D= 50 µm, the effective generation
lifetime must be at least 25 ms. This is a good level of quality that has been already attained in the technology that we
have developed. Further improvement of the technology may be pursued, but the limitation to the maximum active area
compatible with a low DCR level will still be noteworthy. Let us assume to attain an extremely good level in the
fabrication technology, which ensures a very long lifetime
τ
g,eff
=1 s: even with such an outstanding technology, in order
to have DCR not exceeding 1000 c/s it is necessary to limit the maximum diameter of the sensitive area to D = 400 µm.
3. RECENT ADVANCES IN LARGE-AREA SPAD DETECTORS
Fig. 4 shows a schematic cross section of the SPAD structure. A double-epitaxy structure is used in order to reduce
the diffusion effects that adversely affect the time response of the detector [23]. The active n+p junction is built in the
upper low-doped p-epilayer. The buried p+ epilayer provides a low-resistance path to the side ohmic contact. A boron
implantation in the central part of the n+p junction defines the high electric field region, that is, the active area of the
detector. Deep, highly doped n+ regions connected to the bulk n-silicon guarantee electrical isolation between adjacent
SPADs.
In the last years we steadily improved the planar double-epitaxial SPAD technology [24], achieving a reliable fabrication
of high-performance SPAD devices with active area diameter up to 200 µm. The efforts were mainly directed to:
- exploit specific gettering processes, as phosphorus diffusion and p/p+ segregation gettering for removing metal
impurities from the detector active volume, therefore reducing DCR and afterpulsing effects;
Fig. 4 Cross section of the planar SPAD structure.
Proc. of SPIE Vol. 6900 69001D-5
- identify and remove or mitigate all the possible sources of contamination in the detector processing, with special
care to transition metal contamination;
- employ a lower electric field within the p-n junction depletion region to attenuate the band-to band-tunneling and
field-enhanced generation of carriers, thus making it possible to reduce more effectively the DCR by cooling the
detector. In order not to adversely affect the PDE, a reduction of the peak electric field must be compensated by a
suitable increase in the width of the electric field profile.
In the following the performance of the large-area SPAD devices recently fabricated will be reviewed, pointing out the
most significant improvements.
3.1. Photon Detection Efficiency
Fig. 5 shows the photon detection efficiency of a 200 µm SPAD device as a function of wavelength, measured at
different excess bias voltages, V
exc
. At V
exc
= 5 V the PDE has a peak of 52% at 550 nm and it is about 15% at 820 nm
wavelength. These figures are consistent with the total thickness of the epitaxial layer (about 5µm). Similar results have
been obtained with 50 and 100 µm SPAD devices. The PDE is quite uniform all over the wafer: small fluctuations were
observed (std deviation
5% of the average value), mainly due to the thickness fluctuations of the top oxide layer. By
increasing V
exc
to 10 V the peak PDE increases to a remarkable 68%. The PDE increases with V
exc
mainly due to the
increase of the avalanche triggering probability [22]. The small increase of the depletion region thickness gives a
negligible contribution. However, also the DCR increases at higher V
exc
, setting a trade-off that must be carefully
evaluated when seeking the maximum sensitivity.
If the detector noise is dominant, a suitable figure of merit for the sensitivity is the noise equivalent power (NEP),
which is defined as the signal power required to attain a unity signal-to-noise ratio within 1-s integration time:
DCR2
PDE
h
NEP
= (8)
Fig. 6 shows the typical dependence of the NEP on V
exc
for a 50 µm diameter SPAD operating at room temperature. It is
clearly visible that the optimal value of V
exc
is around 4–5 V, where the NEP curves show a broad minimum. Conversely,
if the measurement is background or photon noise limited the optimal V
exc
should be sufficiently high to push the PDE
close to saturation.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
400 500 600 700 800 900 1000
Wavelength (nm)
Photon Detection Efficiency
10 V
7 V
5 V
Excess Bias Voltage
Fig. 5 Photon detection efficiency of a 200 µm SPAD device as a function of wavelength. Measurements were performed at
different excess bias voltages.
Proc. of SPIE Vol. 6900 69001D-6
123456789
Excess bias voltage (V)
Noise Equivalent Power (W Hz )
-1/2
10
-15
10
-16
10
-17
λ
= 850 nm
λ
= 550 nm
Fig. 6 NEP of a 50 µm diameter SPAD as a function of the excess bias voltage. Measurements were performed at room
temperature.
3.2. Dark count rate
Fig. 7 shows the DCR as a function of temperature for SPAD devices having different active area diameters. The
breakdown voltage varies between 33.8 V and 36 V over the temperature range from -50 °C to 20 °C. Measurements
were performed at a constant excess bias voltage of 5 V. The DCR decreases almost exponentially with temperature: at a
relatively low temperature of –25 °C (easily obtainable with a thermoelectric cooler) the typical DCR is 5, 50 and
1500c/s for the 50, 100 and 200 µm SPADs.
0.1
1
10
100
1000
10000
100000
-50 -40 -30 -20 -10 0 10 20
Temperature (°C)
Dark count rate (c/s)
200 µm
50 µm
100 µm
Fig. 7 Dark count rate as a function of temperature for SPAD devices having different active area diameters.
The advantage of an engineered electric field profile is clearly visible in Fig. 8, where the DCR of two 50 µm SPAD
devices made by using different processes are plotted. The DCR of the SPAD device with “standard” electric field [25]
Proc. of SPIE Vol. 6900 69001D-7
can be accurately fitted by taking into account band-to-band tunneling (dotted line) and field-enhanced Shockley-Read-
Hall generation (dashed line): the weights of these two contributions become equivalent at a temperature of about -5 °C
(corner temperature). Since the dependence of tunneling on temperature is relatively weak, just a minor advantage is
gained by cooling the device well below the corner temperature. Conversely, for the SPAD device with “engineered”
electric field no evidence of the onset of tunneling is observed down to -50 °C: cooling can thus be exploited for
obtaining a strong reduction of the DCR. Furthermore, the DCR of devices with engineered electric field profile is lower
also at room temperature, due to the reduced weight of the field-enhanced generation.
0.1
1
10
100
1000
10000
-80 -60 -40 -20 0 20
Temperature (°C)
Dark count rate (c/s)
SPAD with "standard" electric field
SPAD with "engineered" electric field
Fig. 8 Dark count rate as a function of temperature for 50 µm SPAD devices with “standard” and “engineered” electric field profile.
DCR uniformity was assessed by testing about 2000 SPAD devices all over the wafer at room temperature. The
percent point distribution function (i.e. the inverse of the cumulative distribution function) of the DCR is shown in
Fig. 9a for SPAD devices having different active area diameter. It must be noted that the distribution curves for the 100
and 200 µm devices are similar, whereas the DCR distribution for the 50 µm shows a relatively flat region (about 35% of
the 50 µm SPAD devices show a DCR between 1 and 2 kc/s) followed by a steep increase. Fig. 9b shows the DCR
distribution of the most recent SPAD generation. Most of the sources of metal contamination in the fabrication process
were identified and removed or mitigated. As a result, the DCR was reduced by a factor of 5 on the average. It must be
noted that 50% of the 50 µm SPAD devices show a DCR below 500 c/s.
0 20406080100
% of devices with DCR < N
DL
N
DL
: limit to DCR (c/s)
200 µm
100 µm
50 µm
10
10
10
10
10
2
3
4
5
6
Fig. 9 Percentage of SPAD devices (horizontal scale) found within a given limit of the individual dark counting rate (vertical
scale) measured at room temperature. a) previous SPAD generation; b) improved SPAD generation with reduced metal
contamination.
0 20406080100
% of devices with DCR < N
DL
N
DL
: limit to DCR (c/s)
200 µm
100 µm
50 µm
10
10
10
10
10
2
3
4
5
6
a) b)
Proc. of SPIE Vol. 6900 69001D-8
Afterpulsing Probability Density (ns 1)
0 0 0
& 61
0 0 0
C"
0
0
0
0
0
-I
)
01
0
C)
I')
-I
01
0
C)
-U
a
1*.
0)
C"
0
0
Ni
0
0
0
Data shown in Fig. 8 and 9a are consistent with the presence of two types of defects within the depletion region of
the junction [24]. Defects of the first type (type A) have an energy level close to the midgap and a density of about 4
10
8
cm
-3
. The probability of no occurrence of such defects is 40% for a 50 µm device, 3% for a 100 µm device and
practically negligible for a 200 µm device. This observation may justify the relatively flat region in the lower part of the
distribution curve in Fig. 9 for the 50 µm devices. Defects of the second type (type B) have a greater density (between
10
9
and 10
10
cm
-3
), so that they may occur with high probability also in smaller devices. However, their energy level is
displaced by about Eg/4 from the midgap. In general, good devices have no type A defects and their DCR decreases very
steeply by lowering the temperature. Devices with dominant type A defects show a DCR that decreases with a smaller
slope, whereas devices with a mix of non-dominant type A and type B defects show a DCR that decreases with
intermediate slope [24].
3.3. Afterpulsing
Afterpulsing measurements were performed at different temperatures by using the time-correlated carrier counting
(TCCC) technique [26]. TCCC essentially consists in: a) filling the deep levels with a pulsed stimulus; b) measuring the
time interval from the filling pulse to the detection of a released carrier; and c) repeating the procedure for collecting a
histogram of the carrier emissions versus time. Deep levels can be populated by current or light pulses. We adopted an
avalanche current pulse, which has the advantage of filling only the traps involved in the normal device operation. Note
that only electrons released in the p-side and holes released in the n-side cross the high-field region, thus being able to
trigger the avalanche. Moreover, during the avalanche, minority carriers are present only where impact ionizations occur.
It follows that TCCC with avalanche filling stimuli is inherently sensitive only to minority-carrier traps in the high-field
region.
Fig. 10 shows the probability density in time for the occurrence of an afterpulse after an initial avalanche pulse for a
200 µm SPAD device operated at 5 V excess bias voltage. An active quenching circuit (AQC) with a dead time of 80 ns
was used in the experiments. It is worth noting that the probability of having an afterpulse quickly decays, being
negligible after about 1 µs from the initial avalanche pulse. The total afterpulsing probability is typically 2% at room
temperature and it increases up to 6% at -25 °C. This increase is due to the dependence of the trap emission lifetime on
temperature and to the presence of a fixed dead time. At lower temperatures, the emission lifetime of a given trap gets
longer. Accordingly, the probability that a carrier is emitted after the dead time (thus being able to trigger an avalanche)
gets higher.
Fig. 10 Afterpulsing probability density for a 200 µm SPAD device, measured at different temperatures.
Total afterpulsing probabilities ranging from 0.5% to 2% were typically obtained with 50 µm and 100 µm devices
operated at -15 °C with an excess bias voltage of 5 V.
It must be pointed out that it is very difficult to extract significant information about the nature and the
concentration of deep levels from afterpulsing measurements. This is because: i) the electric field has a strong influence
on the emission lifetimes, ii) the extent to which lifetimes are affected depends on the energy and on the charge state of
the deep level, and, iii) the electric field is not uniform in the depletion region.
Proc. of SPIE Vol. 6900 69001D-9
3.4. Time resolution
The time resolution of a SPAD device is determined by the precision with which the arrival instant of the incident
photon on the photodetector is identified. We recently demonstrated [27] that remarkable timing performance is
achievable with large-area SPADs, provided that the avalanche current is sensed at very low level (about a hundred µA
level), when the multiplication process is still confined within a small area around the photon absorption point. In order
to perform a true low-level sensing of the avalanche current it is mandatory to preserve the shape of the first part of the
leading edge by minimizing any filtering action. To this end, we designed and patented [28] a special current pick-up
circuit that can be added to any of the quenching circuit configurations described in the technical and scientific literature.
This circuit employs AC coupling with very short time constant for extracting a signal having a rise-time comparable to
that of the avalanche current. The extracted signal has a very short duration, which makes it possible to maintain very
good timing performance up to very high counting rate.
Time resolution measurements were performed by using an ultra-fast diode (Antel MPL-820 laser module) emitting
15 ps FWHM optical pulses at 820 nm wavelength. Fig. 11 shows the time resolution FWHM of SPAD devices having
different diameters as a function of the threshold level of timing discriminator [27]. Measurements were performed at
5 V excess bias voltage. It is confirmed that a time resolution of 35 ps or slightly less can be achieved regardless of the
device diameter, provided that the threshold is reduced below 10 mV. Fig. 12 shows the time resolution FWHM of a
50 µm-diameter SPAD device operated at different excess bias voltages as a function of the threshold level of timing
discriminator. These measurements clearly show that the time resolution is almost insensitive to the excess bias voltage
provided that the threshold is reduced below 10 mV.
0
50
100
150
200
250
300
350
0 100 200 300 400 500
Threshold voltage (mV)
Time resolution FWHM (ps)
100 µm
50 µm
20 µm
Fig. 11 Time resolution FWHM of SPAD devices having different diameters as a function of the threshold voltage of timing
discriminator. SPADs were operating at 5 V excess bias voltage.
Even more interestingly, we performed the same measurements on 50 µm-diameter SPAD devices belonging to
three different generations. The devices mainly differ in their peak electric field, which decreases by going from
generation #1 to #3. By looking at Fig. 13, it can be concluded that the lower the electric field the steeper is the increase
of the time resolution FWHM with the threshold voltage. Nevertheless, comparable time resolutions FWHM can be
attained by lowering the threshold voltage below 10 mV.
In summary, the use of a low current threshold breaks the complex trade-off between time resolution, active area, and
electric field, thus enabling the fabrication of large-area SPAD devices with excellent DCR and timing performance.
Proc. of SPIE Vol. 6900 69001D-10
0
50
100
150
200
250
300
350
400
0 100 200 300 400 500
Threshold voltage (mV)
Time resolution FWHM (ps)
V
exc
= 5 V
V
exc
= 7 V
V
exc
= 10 V
Fig. 12 Time resolution FWHM of 50µm-diameter SPAD operated at different excess bias voltages as a function of the threshold
voltage of timing discriminator.
Finally, Fig. 14 shows the time response of a 200 µm SPAD detector to a picosecond laser pulse at 820 nm
wavelength. The curve shows a prompt peak with a full-width at half maximum of 35 ps and a clean exponential
diffusion tail with a time constant of 280 ps.
20
40
60
80
100
120
140
160
180
200
220
0 100 200 300 400 500
Threshold voltage (mV)
Time resolution FWHM (ps)
Run #1
Run #2
Run #3
Fig. 13 Time resolution FWHM of 50µm SPAD devices belonging to different generations as a function of the threshold voltage
of timing discriminator.
Proc. of SPIE Vol. 6900 69001D-11
1
10
100
1000
10000
100000
11.5 12.0 12.5 13.0 13.5 14.0 14.5
Time (ns)
Counts
FWHM = 35 ps
Fig. 14 Time resolution curve of a 200 µm SPAD detector. The curve shows a prompt peak with a full-width at half maximum of
35ps and a clean exponential diffusion tail with a time constant of 240ps.
4. CONCLUSIONS
SPAD devices fabricated in planar epitaxial technology offer the typical advantages of microelectronic devices and
provide remarkable performance in single photon counting (SPC) and time-correlated single photon-counting (TCSPC).
Dedicated silicon fabrication technologies provide to the SPAD device designer the quality and flexibility necessary for
attaining further improvements requested by the users. In particular, there is realistic prospect of commercial
development of SPADs with diameter larger than 100 µm, high photon detection efficiency in visible range and excellent
timing resolution. To achieve this objective it is necessary to address both device design and technology issues and
circuit design issues. Our approach to such challenge was a combination of different strategies that is ultra-clean
fabrication process, specific gettering procedures, electric field profile engineering, and low-level detection of the
avalanche current leading edge. By adopting these strategies we fabricated SPAD devices with 200µm active area
diameter. At moderately low temperature (-25 °C with Peltier cooler) these devices have a typical DCR of 1500 c/s and it
is not difficult to select devices with less than 1000 c/s. A fairly low total afterpulsing probability of 2% was measured
with a dead-time of 80 ns. The photon detection efficiency peaks at 52% around 550 nm and stays above 30% over all
the visible range. By using our patented pulse pick-up circuit for processing the avalanche current, a photon timing
resolution
35 ps FWHM was obtained at room temperature, not dependent on the size of the SPAD active area.
These unprecedented characteristics make these SPAD devices a real alternative to MCP-PMTs in demanding TCSPC
applications.
5. ACKNOWLEDGEMENTS
This work was supported by European Commission, Sixth Framework Programme, Information Society Technologies
(NANOSPAD project) and the Italian Ministry of University and Research (MIUR-FIRB project n. RBNE01SLRJ;
MIUR-PRIN project n. 2005027857).
REFERENCES
1. J.R. Lakowicz, “Principles of Fluorescence Spectroscopy”, 3
rd
edition, Springer, Berlin (2006).
2. V. O'Connor and D. Phillips, "Time-correlated Single Photon Counting", Academic Press, London (1984).
Proc. of SPIE Vol. 6900 69001D-12
3. Becker, W., “Advanced Time-Correlated Single Photon Counting Techniques”, Springer, Berlin (2005).
4. R3809U MCP-PMT Data Sheet, Hamamatsu Photonics. Available online at: www.hamamatsu.com.
5. P.P.Webb, R. J. McIntyre, J.Conradi , "Properties of Avalanche Photodiodes," RCA Review, 35, 234-278 (1974).
6. H. Dautet, P. Deschampes, B. Dion, A.D. MacGregor, D. MacSween, R.J. McIntyre, C. Trottier, and P. Webb,
"Photon Counting techniques with silicon avalanche photodiodes," Appl.Opt., 32, 3894-3900 (1993).
7. S.Cova, A.Longoni, and A.Andreoni, "Towards picosecond resolution with single-photon avalanche diodes",
Rev. Sci. Instrum. 52, 408-412 (1981)
8. S. Cova, M. Ghioni, A. Lacaita, C. Samori, and F. Zappa, “Avalanche photodiodes and quenching circuits for
single-photon detection,” Appl.Opt., vol. 35, pp. 1956–1963 (1996).
9. A. Spinelli, A. L. Lacaita, “Physics and Numerical Simulation of Single Photon Avalanche Diodes", IEEE Trans.
Electron Devices, 44, 1931-1943 (1997).
10. SPCM-AQ Data Sheet, PerkinElmer Optoelectronics. Available online at: http://opto.perkinelmer.com
.
11. K. König, , I. Riemann, “High resolution optical tomography of human skin with subcellular resolution and
picosecond time resolution” J. Biom. Opt. 8, 432-439, 2003.
12. H. Yang, G. Luo, P. Karnchanaphanurach, T.-M. Louie, I. Rech, S. Cova, L. Xun, and X. S. Xie, “Protein
conformational dynamics probed by single-molecule electron transfer,” Science, 302, 262–266, (2003).
13. A. Rochas, M. Gani, B. Furrer, P. A. Besse, R. S. Popovic, G. Ribordy and N. Gisin, “Single photon detector
fabricated in a complementary metal–oxide–semiconductor high-voltage technology”, Rev. Sci. Instrum. 74, 3263
(2003).
14. F. Zappa, S. Tisa, A. Gulinatti, A. Gallivanoni, and S. Cova, ``Complete single-photon counting and timing
module in a microchip,'' Opt. Lett. 30, 1327 (2005).
15. S. M. Sze, “Physics of Semiconductor Devices”, pp. 520–527, Wiley, New York (1981).
16. Hurkx, G.A.M.; de Graaff, H.C.; Kloosterman, W.J.; Knuvers, M.P.G., “A new analytical diode model including
tunneling and avalanche breakdown”, IEEE Trans. On Elec. Dev., 39, 2090 - 2098 (1992).
17. Hurkx, G.A.M.; Klaassen, D.B.M.; Knuvers, M.P.G., “A new recombination model for device simulation
including tunneling”, IEEE Trans. On Elec. Dev., 39, 331 – 338 (1992)
18. G. Vincent, A. Chantre, and D. Bois, “Electric field effect on the thermal emission of traps in semiconductor
junctions,” J. App. Phys., 50, 5484-5487 (1979).
19. P. A. Martin, B. G. Streetman, and K. Hess,”Electric field enhanced emission from non-Coulombic traps in
semiconductors,” J. Appl. Phys., 52, 7409-7415 (1981).
20. K. Graff, “Metal Impurities in Silicon-device Fabrication”, 2
nd
edition, Springer-Verlag, Berlin (1995).
21. D. Schroder, Semiconductor Material and Device Characterization”, 3
rd
edition, Wiley, New York (2006).
22. W.O. Oldham, R.R. Samuelson and P.Antognetti, "Triggering Phenomena in Avalanche Diodes," IEEE Trans.
Electron Devices, ED- 19, 1056-1060 (1972).
23. A. Lacaita, M. Ghioni, S. Cova, "Double epitaxy improves single-photon avalanche diode performance", Electron.
Lett., 25, 841-843 (1989).
24. M. Ghioni, A. Gulinatti, P. Maccagnani, I. Rech, S. Cova, “Planar silicon SPADs with 200-µm diameter and 35-ps
photon timing resolution”, Proceedings of SPIE -- Volume 6372, Advanced Photon Counting Techniques,
Wolfgang Becker Editor, 63720R, 0277-786X/06/$15 (2006).
25. A. Gulinatti, I. Rech, P. Maccagnani, M. Ghioni, S. Cova, “Large-area avalanche diodes for picosecond time-
correlated photon counting,” Proceedings of 35th European Solid-State Device Research Conference, ESSDERC
2005, 12-16 Sept. 2005, 355-358 (2005).
26. S.Cova, A.Lacaita and G.Ripamonti, "Trapping phenomena in avalanche photodiodes on nanosecond scale," IEEE
Electron.Dev.Lett., 12, 685-687 (1991).
27. A. Gulinatti, P. Maccagnani, I. Rech, M. Ghioni, and S. Cova, ``35 ps time resolution at room temperature with
large area single photon avalanche diodes,'' Electron. Lett. 41, 272 (2005).
28. S. Cova, M. Ghioni, and F. Zappa, ``Circuit for high precision detection of the time of arrival of photons falling on
single photon avalanche diodes,'' U.S. Patent 6 384 663 B2, May 7, 2002.
Proc. of SPIE Vol. 6900 69001D-13
... On the quenching circuit side, a study showed that the threshold of the discriminator has an impact on the SPAD SPTR [26]. In the current state-of-the art, the best timing jitter is obtained at a low threshold (<100 mV) using a comparator [20,23,27] instead of a classic inverter, but it requires a higher static power consumption. Nevertheless, a study showed promising results with a standard CMOS inverter down to 7.5 ps FWHM for a 25 µm SPAD [28]. ...
... It is also important to validate if the threshold has an impact on the SPAD SPTR. For instance, most bigger SPADs or externally quenched SPADs do not have similar performances in a wide threshold range [23,27]. Furthermore, if the cascode transistor that increases the excess voltage is designed to minimize its impact on the slope and overdrive of the input signal, the cascode does not degrade the timing resolution. ...
Article
Full-text available
In the field of radiation instrumentation, there is a desire to reach a sub-10 ps FWHM timing resolution for applications such as time-of-flight positron emission tomography, time-of-flight positron computed tomography and time-resolved calorimetry. One of the key parts of the detection chain for these applications is a single-photon detector and, in recent years, the first single-photon avalanche diode (SPAD) with a sub-10 ps timing resolution was presented. To reach such a timing resolution, the SPAD was read out by an operational amplifier operated in open-loop as a comparator. This paper presents a comparison between comparators and inverters to determine which type of leading-edge discriminator can obtain the best single-photon timing resolution. Six different quenching circuits (QCs) implemented in TSMC 65 nm are tested with SPADs of the same architecture and in the same operation conditions. This allows us to compare experimental results between the different QCs. This paper also presents a method to measure the SPAD signal slope, the SPAD excess voltage variation and simulations to determine the added jitter of different leading-edge discriminators. For some discriminator architectures, a cascode transistor was required to increase the maximum excess voltage of the QC. This paper also presents the impact on the single-photon timing resolution of adding a cascode transistor for a comparator or an inverter-based discriminator. This paper reports a 6.3 ps FWHM SPTR for a SPAD read out by a low-threshold comparator and a 6.8 ps FWHM SPTR for an optimized 1 V inverter using a cascode transistor for a higher excess voltage.
... Sur la Figure 1.9, on observe que la probabilité d'afterpulsing augmente avec le diminution de la température. En effet, à basses températures, le temps de rétention des porteurs est plus grand ; des avalanches peuvent se produire une fois le temps mort (correspondant au temps d'étouffement et de recharge de la SPAD) écoulé [26]. De plus, pour réaliser des mesures optimales, il est préférable d'avoir un temps de recharge court et donc de privilégier l'utilisation d'un circuit d'étouffement actif plutôt que passif. ...
... Par ailleurs, l'intensité du courant, qui varie avec la tension d'excès, engendre aussi une augmentation de la probabilité d'afterpulsing [27]- [29]. . FIGURE 1.9 -Exemple de mesures de la probabilité d'afterpulsing sur une SPAD silicium de diamètre 200 µm [26] ...
Thesis
L'objectif de cette thèse concerne la simulation, la conception et la caractérisation de nouvelles structures de diodes à avalanche à photon unique (Single Photon Avalanche Diode - SPAD) implémentées dans la technologie CMOS FD-SOI (Fully Depleted Silicon On Insulator) 28nm de STMicroelectronics. Les photodétecteurs SPAD présentent une grande sensibilité de détection (associée à un temps de réponse très court) qui fait d’eux d’excellents candidats pour la mesure du temps de vol (Time Of Flight – ToF) dans des applications de télémétrie, de reconnaissance faciale et de LIDAR (Light Detection And Ranging) pour les voitures autonomes. L’intégration de la SPAD en CMOS FD-SOI permet de créer un pixel intrinsèquement 3D, i) en incorporant la SPAD au niveau de la jonction PW (P-Well) / DNW (Deep N-Well) dans le silicium bulk sous l’oxyde enterré (BOX) et ii) en utilisant le film silicium situé au-dessus du BOX pour intégrer l'électronique associée au détecteur (circuits d'étouffement et d'adressage), tout en optimisant le facteur de remplissage avec une approche BSI (back side illumination). Les SPAD réalisées dans la technologie native (avec respect des règles de dessin) ont mis en évidence plusieurs points faibles : un DCR (Dark Count Rate) élevé pour des tensions d'excès faibles (500Hz/µm2 à Vex = 0.5V pour une tension de claquage de 9.5V) ainsi qu'un claquage prédominant sur la périphérie de la zone active. Dans ce contexte, les travaux présentés dans cette thèse ont porté sur l'optimisation des performances électriques de la SPAD FD-SOI par des modifications de la structure respectant ou non le procédé de fabrication : adaptation des conditions d’implantation du caisson profond DNW, remaniement des tranchées STI (Shallow Trench Isolation) etc. Les structures SPAD-FD-SOI ainsi optimisées ont démontré expérimentalement un bien meilleur niveau de DCR (17Hz/µm2 à Vex = 1V pour une tension de claquage de 15.8V). Des caractérisations électro-optiques préliminaires ont été réalisées avec une probabilité de détection des photons de l’ordre de 7% à Vex = 1V et une longueur d’onde de 650nm. Même si ces travaux n’ont pas permis d’atteindre les performances des SPAD les plus performantes de l’état de l’art, ils ont exploré de nombreuses voies d’optimisation, certaines conduisant à une amélioration significative des performances des SPAD réalisées dans cette technologie. La poursuite de ces travaux (association de ces structures SPAD FD-SOI optimisées avec une électronique intégrée performante, amincissement des dispositifs pour opérer avec un éclairage par la face arrière etc.) devrait permettre de réaliser des pixels SPAD intrinsèquement 3D (sans recours à du collage de wafers) très performants dans le proche infrarouge pour les applications d’imagerie 3D embarquées.
... Although the active quenching circuit (AQC) and variable load quenching circuit (VLQC) can effectively improve the speed of largesize GmAPD, they requires a rather complex quenching circuit. [17][18][19] Currently, the active-area diameter of discrete GmAPD s is primarily tens of micrometers, 20 with the largest reaching around hundreds of micrometers, 15,20,21 and larger ones (i.e. mm scale) are mostly realized on the basis of detector array which usually faces fill factor difficulties. ...
Article
While the larger photosensitive area of Geiger-mode Avalanche Photodiode (GmAPDs) enhances their detection range and signal collection, improving their utility in weak light detection, their practicality is limited by long recovery times, high afterpulsing probability (AP), and excessive jitter. Utilizing a Dynamic Memristor (DM) as a quenching resistor, this research improves the count rate of a large-size GmAPD by 100× at an overvoltage of 2.5 V, compared to a fixed resistor (FR)-quenched GmAPD. Furthermore, at 1 MHz photon pulse frequency, jitter time is reduced from 3.60 ns to 0.48 ns, and afterpulsing probability is effectively mitigated from 30.88% to 8.58%.
... The typical structure of a double epitaxial thin SPAD developed at Politecnico di Milano [33] is shown in Fig. 1(a). The device is fabricated starting from an n-type substrate on top of which has been grown a p-type epitaxial layer, composed by a p+ buried layer and a p-quasi-intrinsic layer. ...
Article
Full-text available
Predictive and reliable models are key tools for the development of the next generations of single photon avalanche diodes (SPADs). Models are indeed crucial to evaluate the performance of prospective detector structures and to down select the most promising solutions before developing and running a dedicated fabrication process. To ensure predictability, models must be extensively validated against experimental data. In particular, the model must be applied to existing detectors and the results obtained from simulations must be compared with the measurements performed on the same detectors. The ability to accurately extract the doping profile along the SPAD active region plays a crucial role in the validation flow, because SPAD properties generally exhibit a strong dependence on the electric field. In this article we will discuss why widely-adopted doping profile techniques do not allow the calculation of the electric field with an accuracy sufficient for predictive SPAD modeling. Then, we will present a technique we developed to extract accurate doping profiles starting from an approximate device-structure built with process simulations and/or secondary ion mass spectroscopy (SIMS) measurements. The technique combines capacitance-vs-voltage measurements with device simulations to implement an inverse modeling scheme. Finally, we will show how the proposed technique allows us to accurately reproduce the breakdown voltage of a large set of SPADs.
... SPAD high sensitivity with low photon flux detection capabilities [2], excellent jitter or time resolution [3], and small dead time [4] make them suitable for future space-borne Lidar missions. CMOS technologies development improved these devices, with the possibility to combine the SPAD detector and its driving electronic [5] on the same integrated circuit. ...
Article
Full-text available
Single-Photon Avalanche Diodes (SPAD) in Complementary Metal-Oxide Semiconductor (CMOS) technology are potential candidates for future “Light Detection and Ranging” (Lidar) space systems. Among the SPAD performance parameters, the Photon Detection Probability (PDP) is one of the principal parameters. Indeed, this parameter is used to evaluate the SPAD sensitivity, which directly affects the laser power or the telescope diameter of space-borne Lidars. In this work, we developed a model and a simulation method to predict accurately the PDP of CMOS SPAD, based on a combination of measurements to acquire the CMOS process doping profile, Technology Computer-Aided Design (TCAD) simulations, and a Matlab routine. We compare our simulation results with a SPAD designed and processed in CMOS 180 nm technology. Our results show good agreement between PDP predictions and measurements, with a mean error around 18.5%, for wavelength between 450 and 950 nm and for a typical range of excess voltages between 15 and 30% of the breakdown voltage. Due to our SPAD architecture, the high field region is not entirely insulated from the substrate, a comparison between simulations performed with and without the substrate contribution indicates that PDP can be simulated without this latter with a moderate loss of precision, around 4.5 percentage points.
Article
Full-text available
Picosecond timing of single photons has laid the foundation of a great variety of applications, from life sciences to quantum communication, thanks to the combination of ultimate sensitivity with a bandwidth that cannot be reached by analog recording techniques. Nowadays, more and more applications could still be enabled or advanced by progress in the available instrumentation, resulting in a steadily increasing research interest in this field. In this scenario, single-photon avalanche diodes (SPADs) have gained a key position, thanks to the remarkable precision they are able to provide, along with other key advantages like ruggedness, compactness, large signal amplitude, and room temperature operation, which neatly distinguish them from other solutions like superconducting nanowire single-photon detectors and silicon photomultipliers. With this work, we aim at filling a gap in the literature by providing a thorough discussion of the main design rules and tradeoffs for silicon SPADs and the electronics employed along them to achieve high timing precision. In the end, we conclude with our outlook on the future by summarizing new routes that could benefit from present and prospective timing features of silicon SPADs.
Article
The silicon photomultipliers (SiPMs) emerged as a promising solution in many applications, like high-energy physics experiments and recently, they are the detector of choice for the readout of liquid noble gases scintillators (e,g, liquid Xenon and Argon) in very-large-area physics experiments. Here the SiPMs are operated at cryogenic temperatures. Important studies have been done to optimize SiPM performance for such conditions and we developed the NUV-HD-cryo and the VUV-HD-cryo technologies, with sensitivity optimized for the blue-wavelength and NUV range, or for the VUV range respectively. Important technological improvements have been demonstrated: (i) reduction of electric field, to reduce band-to-band tunneling, (ii) doping profiles modifications to reduce afterpulsing at low temperatures and (iii) reduction of quenching resistor variation over temperature. In this work we show and compare the recent characterization results of primary noise, correlated noise. We also show the dependences of the photon detection efficiency (PDE) and photon-number resolution of these SiPMs over temperature, between 300 K and 75 K. Primary dark count rate reduces to few counts per second already at 200 K, and of 7 orders of magnitude going toward liquid nitrogen temperature, afterpulsing increment is mitigated, being less than 15%. The PDE in the blue-wavelength region remain high in the investigated temperature range, while it changes at longer wavelengths and good photon number resolution is preserved.
Article
Lidar has already achieved long-distance and high-accuracy detection due to the improved sensitivity and timing accuracy of single-photon detectors (SPDs). However, single-photon Lidar systems based on semiconductor detectors are usually susceptible to dark counts and afterpulsing. Recently, superconducting nanowire single-photon detectors (SNSPDs) have shown promise toward single-photon Lidar systems owing to their high detection efficiency, high maximum count rate, low dark count rate and low timing jitter. An in-depth understanding of SNSPD Lidar can promote its further rational development and application. This article reviews the advantages of SNSPDs in single-photon Lidar systems and corresponding developments and introduces and prospects its related applications. In the future, the development of SNSPD working bands and array sizes may promote the next generation of single-photon Lidar systems with a longer detection range, higher depth accuracy and shorter acquisition time.
Book
This monograph describes the transition-metal impurities generated during silicon sample and device fabrication. The different mechanisms responsible for contamination are discussed, and a survey is given of their impact on device performance. The specific properties of main and rare impurities in silicon are examined, as well as the detection methods and requirements in modern technology. Finally, impurity gettering is studied along with modern techniques to determine gettering efficiency.
Article
The basic properties of avalanche photodiodes are reviewed and the conditions under which they can be used to best advantage are discussed. Expressions are presented for calculating the avalanche gain, signal-to-noise ratio, detector excess noise factor, effective noise factor under conditions of mixed injection, diode linearity and saturation, noise equivalent power, quantum efficiency, impulse time response, and frequency response. Several commercially available avalanche photodiode structures are compared, and their virtues and disadvantages are discussed. The properties of a reach-through avalanche photodiode are discussed in detail, and experimental results are presented for single-element avalanche diodes with sensitive areas up to 20 mm**2, and quadrant avalanche diodes with 5 mm**2 sensitive areas. Finally, an equivalent circuit for the reach-through avalanche diode is derived.
Book
Optical Signal Recording.- Overview of Photon Counting Techniques.- Multidimensional TCSPC Techniques.- Building Blocks of Advanced TCSPC Devices.- Application of Modern TCSPC Techniques.- Detectors for Photon Counting.- Practice of TCSPC Experiments.- Final Remarks.- References.
Article
Semiconductor Material and Device Characterizationis the only book on the market devoted to the characterization techniques used by the modern semiconductor industry to measure diverse semiconductor materials and devices. It covers the full range of electrical and optical characterization methods while thoroughly treating the more specialized chemical and physical techniques. In the third edition, Professor Schroder has rewritten parts of each chapter and added two new chapters (Charge Based Measurements and Failure Analysis and Reliability), redrawn and updated most figures, and included new problems and approximately 100 new references. * New end of chapter problems * Outdated figures have been redone and replaced with current data * Up-to-date bibliography with over 1400 references * Professor Schroder is recognized as the authority in the field of semiconductor characterization
Article
Scitation is the online home of leading journals and conference proceedings from AIP Publishing and AIP Member Societies
Article
Time-correlated single photon counting (TCSPC) is exploited in emerging scientific applications in life sciences, such as single molecule spectroscopy, DNA sequencing, fluorescent lifetime imaging. Detectors with wide active area (diameter > 100 mum) are desirable for attaining good photon collection efficiency without requiring complex and time-consuming optical alignment and focusing procedures. Fiber pigtailing of the detector, often employed for having a more flexible optical system, is also obtained more simply and with greater coupling efficiency for wide-area detectors. TCSPC, however, demands to detectors also high photon-timing resolution besides low noise and high quantum efficiency. Particularly tight requirements are set for single-molecule fluorescence analysis, where components with lifetimes of tens of picoseconds are often met. Small photon timing jitter and wide area are considered conflicting requirements for the detector. We developed an improved planar silicon technology for overcoming the problem and providing a solid-state alternative to MCP-PMTs in demanding TCSPC applications. We fabricated Single Photon Avalanche Diodes (SPADs) with 200 mum active area diameter and fairly low dark counting rate (DCR). At moderately low temperature (-25 °C with Peltier cooler) the typical DCR is 1500 c/s and it is not difficult to select devices with less than 1000 c/s. The photon detection efficiency peaks at 48% around 530 nm and stays above 30% over all the visible range. A photon timing resolution of 35 ps FWHM (full width at half maximum) is obtained by using our patented pulse pick-up for processing the avalanche current.