Quantum cascade detector utilizing the diagonal

Quantum cascade detector
utilizing the diagonal-transition scheme
for high quality cavities
P. Reininger,∗ B. Schwarz, R. Gansch, H. Detz, D. MacFarland,
T. Zederbauer, A. M. Andrews, W. Schrenk, and G. Strasser
Institute for Solid State Electronics and Center for Micro- and Nanostructures, Vienna
University of Technology, Floragasse 7, Vienna 1040, Austria
∗ [email protected]
Abstract:
A diagonal optically active transition in a quantum cascade
detector is introduced as optimization parameter to obtain quality factor
matching between a photodetector and a cavity. A more diagonal transition
yields both higher extraction efficiency and lower noise, while the reduction
of the absorption strength is compensated by the resonant cavity. The
theoretical limits of such a scheme are obtained, and the impact of losses
and cavity processing variations are evaluated. By optimizing the quantum
design for a high quality cavity, a specific detectivity of 109 Jones can be
calculated for λ = 8 µm and T = 300 K.
© 2015 Optical Society of America
OCIS codes: (040.5350) Photovoltaic; (040.5160) Photodetectors; (040.3060) Infrared;
(040.5570) Quantum detectors.
References and links
1. J. Faist, F. Capasso, D. L. Sivco, C. Sirtori, A. L. Hutchinson, and A. Y. Cho, “Quantum cascade laser,” Science
264, 553 (1994).
2. B. S. Williams, “ Terahertz quantum-cascade lasers,” Nature photonics 1 (9), 517-525 (2007).
3. M. S. Vitiello, G. Scamarcio, V. Spagnolo, S. S. Dhillon, and C. Sirtori, “Terahertz quantum cascade lasers with
large wall-plug efficiency,” Appl. Phys. Lett. 90, 191115 (2007).
4. M. Razeghi, S. Slivken, Y. B. Bai, B. Gokden, and S. R. Darvish, “High power quantum cascade lasers,” New J.
Phys. 11 (12), 125017 (2009).
5. D. Hofstetter, M. Beck, and J. Faist, “Quantum-cascade-laser structures as photodetectors,” Appl. Phys. Lett. 81,
2683 (2002).
6. H. Schneider, C. Sch¨onbein, G. Bihlmann, P. Van Son, and H. Sigg, “High-speed infrared detection by uncooled
photovoltaic quantum well infrared photodetectors,” Appl. Phys. Lett. 70, 1602 (1997).
7. H. Schneider, H. C. Liu, Quantum Well Infrared Photodetectors: Physics and Applications (Springer,2007).
8. M. Graf, G. Scalari, D. Hofstetter, J. Faist, H. Beere, E. Linfield, D. Ritchie, and G. Davies, “Terahertz range
quantum well infrared photodetector,” Appl. Phys. Lett. 84, 475 (2004).
9. M. Graf, N. Hoyler, M. Giovannini, J. Faist, and D. Hofstetter, “InP-based quantum cascade detectors in the
mid-infrared,” Appl. Phys. Lett. 88, 241118 (2006).
10. L. Gendron, M. Carras, A. Huynh, V. Ortiz, C. Koeniguer, and V. Berger, “Quantum cascade photodetector,”
Appl. Phys. Lett. 85, 2824 (2004).
11. L. Gendron, C. Koeniguer, V. Berger, and X. Marcadet, “High resistance narrow band quantum cascade photodetectors,” Appl. Phys. Lett. 86, 121116 (2005).
12. H. C. Liu, C. Y. Song, A. J. SpringThorpe and J. C. Cao, “Terahertz quantum-well photodetector,” Appl. Phys.
Lett. 84, 4068 (2004).
13. F. R. Giorgetta, E. Baumann, M. Graf, Q. Yang, C. Manz, K. K¨ohler, H. E. Beere, D. A. Ritchie, E. Linfield, A.
G. Davies, Y. Fedoryshyn, H. J¨ackel, M. Fischer, J. Faist, and D. Hofstetter, “Quantum cascade detectors,” IEEE
J. Quantum Electron. 45, 1029-1043 (2009).
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6283
14. S. Sakr, E. Giraud, A. Dussaigne, M. Tchernycheva, N. Grandjean, and F. H. Julien, “Two-color GaN/AlGaN
quantum cascade detector at short infrared wavelengths of 1 and 1.7 µm,” Appl. Phys. Lett. 100, 181103 (2012).
15. A. P. Ravikumar, T. A. Garcia, J. De Jesus, M. C. Tamargo and C. F. Gmachl, “High detectivity short-wavelength
II-VI quantum cascade detector,” Appl. Phys. Lett. 105, 061113 (2014).
16. P. Reininger, B. Schwarz, A. Harrer, T. Zederbauer, H. Detz, A. M. Andrews, R. Gansch, W. Schrenk, and G.
Strasser, “Photonic crystal slab quantum cascade detector,” Appl. Phys. Lett. 103, 241103 (2013).
17. A. Harrer, B. Schwarz, R. Gansch, P. Reininger, H. Detz, T. Zederbauer, A. M. Andrews, W. Schrenk, and G.
Strasser, “Plasmonic lens enhanced mid-infrared quantum cascade detector,” Appl. Phys. Lett. 105, 171112
(2013).
18. Y. N. Chen, Y. Todorov, B. Askenazi, A. Vasanelli, G. Biasol, R. Colombelli, and C. Sirtori, “Antenna-coupled
microcavities for enhanced infrared photo-detection,” Appl. Phys. Lett. 104, 031113 (2014).
19. S. Kalchmair, H. Detz, G. Cole, A. M. Andrews, P. Klang, M. Nobile, R. Gansch, C. Ostermaier, W. Schrenk, and
G. Strasser, “Photonic crystal slab quantum well infrared photodetector,” Appl. Phys. Lett. 98, 011105 (2011).
20. S. Kalchmair, R. Gansch, S. Ahn, A. M. Andrews, H. Detz, T. Zederbauer, E. Mujagi´c, P. Reininger, G. Lasser, W.
Schrenk, and G. Strasser, “Detectivity enhancement in quantum well infrared photodetectors utilizing a photonic
crystal slab resonator,” Opt. Express 20, 5622 (2012).
21. P. Reininger, B. Schwarz, H. Detz, D. MacFarland, T. Zederbauer, A. M. Andrews, W. Schrenk, and G. Strasser,
“Diagonal-transition quantum cascade detector,” Appl. Phys. Lett. 105, 091108 (2014).
22. O. Baumgartner, Z. Stanojevic, K. Schnass, M. Karner and H. Kosina, “VSP-a quantum-electronic simulation
framework,” J. Comput. Electron. 12, 701 - 721(2013).
23. A. Delga, M. Carras, L. Doyennette, V. Trinit´e, A. Nedelcu and V. Berger, “Predictive circuit model for noise in
quantum cascade detectors,” Appl. Phys. Lett. 99, 252106 (2011).
24. S. Fan, W. Suh and J. D. Joannoploulos, “Temporal coupled-mode theory for the Fano resonance in optical
resonators,” J. Opt. Soc. Am. A 20, 569-572 (2003).
25. A. F. Oskooi, D. Roundy, M. Ibanescu, P. Bermel, J. D. Joannoploulos, S. G. Johnson, “MEEP: A flexible freesoftware package for electromagnetic simulations by the FDTD method,” Comp. Phys. Comm. 181(3), 687-702
(2010)
1.
Introduction
Quantum cascade detectors (QCD) originated from the concept of quantum cascade lasers
(QCL) [1–4]. It was found, that, by not biasing a QCL, it can be operated as a photovoltaic
detector [5]. At this time, photovoltaic quantum well infrared photodetectors had already been
demonstrated and showed very promising performance [6,7]. By specifically designing a quantum cascade structure for detection, rather than lasing, a significant performance increase compared to unbiased QCLs could be achieved, which established the topic of QCDs [8–11]. To
date, detection has been shown in a broad spectral range, namely from the near-infrared to
the THz, but only in the near-infrared and mid-infrared has room-temperature operation with
decent device performance been demonstrated [12–15]. Because of their versatility, QCDs are
perfectly suited to use in combination with sophisticated cavities and have been demonstrated
to yield performance increase with photonic crystals or plasmonic antennas [16–18]. The highest possible performance is obtained when the quality factor (Q-factor) of the cavity and the
detector material match. So far, this has been achieved by reducing the doping of the detector
and thereby increasing its Q-factor [19,20]. By that, the noise was reduced significantly, but the
responsivity remained unchanged.
An alternative way to adjust Q-factors was introduced with a new QCD design concept, the
diagonal-transition QCD [21], that showed a performance increase of almost an order of magnitude compared to other QCDs at the same wavelength. Here, the optically active transition
occurs between two energy levels that are localized in adjacent wells. The dipole matrix element
can be controlled, to match the cavity Q-factor, by changing the thickness of the separating barrier. A thicker barrier reduces the absorption strength , while the extraction efficiency and the
resistance of the device is increased. In this way, both the responsivity and the detectivity can
be increased. We will show that by fully utilizing both reduced doping and the diagonality of
the optically active transition, the detector performance can be significantly enhanced.
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6284
2.
Quantum design
The performance of photodetectors is commonly quantified through the responsivity or specific
detectivity as figure of merit. The responsivity R is a measure of the unit of electrical output per
optical input power. In the case of QCDs, it can be written as
R=
λ q η pe
,
hc N
(1)
where λ is the wavelength, q the elementary charge, h the Planck constant, c the vacuum speed
of light, η the absorption efficiency, pe the extraction efficiency and N the number of QCD
cascades. The specific detectivity D∗ also contains a noise term and describes the signal to
noise behavior. In the Johnson noise limit, the specific detectivity for QCDs near zero bias can
be expressed as
R0 A
∗
,
(2)
Dj = R
4kB T
where R is the responsivity, R0 is the device resistance, A the device area, kB the Boltzmann
constant and T the temperature. By combining the QCD with a resonant cavity, the lifetime of
photons in the QCD active region is increased, which increases the absorption efficiency η . This
so-called resonant absorption enhancement is only efficient if the cavity and the photodetector
material have similar Q-factors. The Q-factor is defined as
Q = 2π
Energy stored
Energy dissipated per cycle
(3)
and can be linked to the absorption coefficient of a material by
Q=
n
2π n
=
2κ
λα
(4)
where n˜ = n + iκ is the refractive index, λ the wavelength and α the absorption coefficient
related to the entire QCD period. It is assumed that the entire cavity is filled by the absorbing
QCD material. Usually, both quantum-well infrared photodetectors and QCDs have an absorption coefficient that corresponds to a significantly lower Q-factor than that of resonant cavities.
Typical values of α are 1000 – 2000 cm−1 . Using above equation, this corresponds to a detector
Q-factor of approximately 25 – 12.5 in a InGaAs/InAlAs heterostructure, at λ = 8 µm. Such
high absorption is necessary to get reasonable absorption efficiencies in typically used doublepass mesa structures. A higher detector Q-factor (or equivalently, lower detector absorption)
can be obtained by the reduction of the doping. It was shown that the specific detectivity could
be significantly increased by reducing the doping of a quantum well infrared photodetector and
combining it with a photonic crystal slab resonant cavity [20]. The benefit of lower doping
is the reduction of the noise current density. In QCDs, a lower doping density increases the
resistance, but√according to Eq. 2, an c times higher resistance increases the specific detectivity by only c. With the diagonal-transition QCD, another means to control the absorption
coefficient has been introduced, namely the diagonality of the optically active transition. The
bandstructure of the diagonal-transition QCD that has been used for this analysis is depicted in
figure 1. The material system is InGaAs/InAlAs lattice matched to InP. The layer thicknesses
in nm are 4.4/5.4/1.85/2.2/4.4/2.6/4.4/2.8/ 4.4/2.954.6/3.14.5/2.2/4.55/3.4/4.7/3.6/4.9/3.8. The
Barriers are indicated in bold. In grown structures, the underlined layers are doped. The general
conclusions are also valid for other QCD designs, where the optically active transition occurs
between two adjacent wells. The optical transition is indicated by the black arrows. By controlling the thickness of the barrier between the two active energy levels (indicated by the shaded
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6285
Energy (eV)
0.6
0.5
0.4
0.3
0.2
0
20
40
60
80
100
Direction of growth (nm)
Fig. 1. Bandstructure of the diagonal-transition QCD that is used for this analysis. It is
based on a diagonal-transition scheme, thus the resistance, the extraction efficiency and
the absorption can be controlled by the thickness of the barrier between the active wells,
indicated by the shaded green region.
green region), the overlap integral and the dipole matrix element can be tuned. A thicker barrier
increases the extraction efficiency and gives a higher electrical resistance [21]. This yields a
performance improvement despite the lower absorption coefficient of the diagonal-transition
QCD, even without a resonant cavity. In contrast to the doping density, both the responsivity and the specific detectivity are directly proportional to the extraction efficiency. A number of simulations for varying doping density and barrier thickness was performed, using a
single-particle Monte-Carlo transport simulator for quantum cascade devices [22]. The following transport mechanisms were included: acoustic deformation potential, optical deformation
potential, polar optical phonons, alloy scattering and interface roughness scattering. Using the
obtained scattering rates, the resistance, the responsivity and the specific detectivity was calculated. The resistance was calculated based on the model from [23]. A variation of the doping
density generates a change of the bandstructure through band-bending, that can be compensated
by a careful quantum design. This can be done in the range of sheet doping density used in this
analysis. At even higher doping densities the relationship between extraction efficiency, resistance and sheet doping is strongly non-linear and the figures of merit deteriorate. That is why
we limited the sheet doping density to n2D = 1012 cm−1 . We chose to perform the simulations
without self-consistency for rapid prototyping of designs. Of course, for the design of growth
structures, self-consistent simulations are necessary.
In Fig. 2, the absorption coefficient α , the extraction efficiency pe and the resistance R0 of
the QCD are depicted as a function of the doping density n2D and the barrier thickness d. The
range of the barrier thicknesses was from 1-5 nm. With this we obtained dipole matrix elements
of the optically active transition of 2.67-0.45 nm. Arbitrary values of α can be reached by a
variety of combinations of n2D and d. Each different set of n2D and d for a given α has a
different extraction efficiency and resistance. Generally, the resistance and extraction efficiency
are higher for designs with lower absorption.
3.
Quality factor optimization
To find the optimal performance in terms of either responsivity or specific detectivity we have
to formulate an equation that describes the dependence of the absorption efficiency η on the
absorption coefficient α of the QCD, which is valid regardless of the used cavity. To obtain the
results for a particular example, one needs to obtain the different Q-factors that describe the
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6286
Sheet doping density (cm−2 )
(a)
10
Absorption coefficient (cm−1 )
12
(b)
(c)
Extraction efficiency
104
Resistance (Ohm)
106
1
0.9
103
105
0.8
0.7
102
1011
104
0.6
101
103
0.5
0.4
100
102
0.3
1010
1 1.5 2 2.5 3 3.5 4 4.5
Barrier thickness (nm)
10−1
1 1.5 2 2.5 3 3.5 4 4.5
101
0.2
1 1.5 2 2.5 3 3.5 4 4.5
Barrier thickness (nm)
Barrier thickness (nm)
Fig. 2. Absorption coefficient α , extraction efficiency pe and resistance R0 as twodimensional plots versus the thickness of the active barrier and the sheet doping density
at T = 300 K. Standard QCDs have a typical absorption coefficient in the range of 1000 –
2000 cm−1 . Such high absorption is obtained for either a low barrier thickness or a high
doping density (or a combination of both), yielding low extraction efficiency and low resistance. High extraction efficiency and high resistance can only be obtained for devices with
a low absorption coefficient.
cavity and use it in the equation. The temporal coupled mode theory [24] is a powerful tool that
can be used to evaluate such a problem. We obtain the following equation for the absorption
efficiency:
2 Qc1Q
d
η= (5)
2 ,
2 1
ω0
4 ω − 1 + Qc + Q1 + Q1
d
l
where Qc is the so-called cavity Q-factor that describes the coupling of the resonator mode to
freespace radiation, Qd the detector Q-factor that describes the intersubband absorption strength
of the QCD material, Ql the loss Q-factor describing the total loss of the device and ω0 the
resonance frequency of the cavity mode. The spectral width of a mode with a given Q-factor can
be obtained using ωFW HM = ωres /Q. We assume in this analysis a constant frequency ω = ω0 ,
such that the frequency dependent term in Eq. 5 vanishes. With this equation we evaluate the
impact of variations of the different Q-factors on the performance of the device. To obtain the
behavior of the resonant cavity photodetector under ideal circumstances, we initially neglect all
losses and set Q1 = 0. We can calculate the responsivity and specific detectivity as a function of
l
n2D and d, by combining Eq. (5) with Eqs. (1) and (2) and applying it to the values from Fig. 2.
Figure 3 shows the responsivity and detectivity as a function of n2D and d for a coupling
strength of Qc = 200. The maximum of the responsivity for each fixed doping density occurs at
the critical coupling condition, i.e. when η is maximum or, equivalently, Qc = Qd . If we look
at the responsivity at a higher doping density, its maximum occurs at a higher barrier thickness,
because a thicker barrier is necessary to maintain the same detector absorption. However, the
maximum of the specific detectivity does not occur at the critical coupling condition. For a
fixed doping density the maximum of the detectivity is at thicker barriers than the responsivity.
The reason is the inclusion of the resistance in its definition, which is larger for thicker barriers.
Generally formulated, the maximum of the specific detectivity occurs at a lower value of Qd
than Qc . For different values of Qc , we will find the maxima at different values of n2D and
d, or equivalently at a different Qd . Also, the magnitude of the overall maxima will change
depending on Qc .
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6287
(a)
Sheet doping density n2D (cm−2 )
Detectivity Dj∗ (Jones)
(b)
Responsivity R (mA/W)
1012
2.5x108
80
70
2.0x108
60
50
10
11
1.5x108
40
1.0x108
30
20
10
Qc = 200
10
0.5x108
10
Qc = 200
0
0
1
1.5
2
2.5
3
3.5
4
4.5
1
1.5
Barrier thickness d (nm)
2
2.5
3
3.5
4
4.5
Barrier thickness d (nm)
Fig. 3. Calculated responsivity and Detectivity at T = 300 K versus the barrier thickness and
sheet doping density. All losses were neglected and Q1l was set to zero. The maximum of
the responsivity occurs for the so-called critical coupling condition, i.e. when the absorption efficiency is maximal or equivalently when Qd = Qc . The maximum of the specific
detectivity occurs at a lower Qd .
100
1.50x109
T = 300K
1.25x109
1.00x109
90
0.75x109
80
0.50x109
70
60
0.25x109
100
1000
Optimum Dj∗ (Jones)
Optimum R (mA/W)
110
0
Cavity Q factor Qc
Fig. 4. Optimum responsivity and specific detectivity versus the cavity Q-factor, where the
losses are neglected. For each value of Qc , Qd (n2D , d) was optimized for R and then again
for D∗j . The optimum of the responsivity occurs at the critical coupling condition, when
Qc = Qd . The steep initial increase of the responsivity at low Qc comes from the increasing extraction efficiency. At higher values of Qc , it converges to the maximal achievable
= 8 µm and for the chosen number of 30 periods. The specific
responsivity for η = 0.5 at λ √
detectivity has an additional R0 term in its definition. For higher detector Q-factors Qd ,
the resistance is also increasing. That means, without losses, there is no theoretical limit
for the specific detectivity.
The behavior of the maximum responsivity and the maximum specific detectivity versus Qc ,
under ideal conditions, is shown in Fig. 4. Both show an increase for higher values of Qc . The
responsivity has a steep initial increase, because in this region the transition gets more diagonal
and the extraction efficiency rises. For even higher Q-factors, the extraction efficiency saturates,
and thus the responsivity converges towards the maximum responsivity for this wavelength
and for a chosen number of 30 periods. The specific detectivity is directly proportional to the
responsivity and the square root of the resistance. We have seen that for high values of Qc , and
thus high values of Qd , the responsivity saturates. However, since the resistance increases with
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6288
Qd , the specific detectivity gets arbitrarily large.
Optimum η
0.5
Ql → ∞, α = 0.0 cm−1
0.4
Ql ≈ 10000, α = 2.5 cm−1
0.3
Ql ≈ 5000, α = 5.0 cm−1
Ql ≈ 2500, α = 10 cm−1
0.2
Ql ≈ 1200, α = 20 cm−1
Ql ≈ 600, α = 40 cm−1
0.1
0
100
1000
10000
Cavity Q factor Qc
Fig. 5. Optimum absorption efficiency versus Qc for different values of waveguide losses.
For each value of Qc and given Ql , Qd was optimized for η . The optimum occurs at the
critical coupling condition, which changes to 1/Qd = 1/Qc + 1/Ql , if losses are taken into
account. An increasing loss imposes an upper limit for reasonable detector Q-factors.
In reality, losses of different origins impose an upper limit to reasonable Q-factors. The most
dominant loss mechanisms originate from the finite dimensions of the cavity (often referred to
as the horizontal Q-factor) and free carrier absorption in the waveguide. The impact of loss in
the system is evaluated by a combined loss Q-factor, that is obtained by using the following
summation
1
1
=∑
, i ∈ {1, 2, 3, ..., N}
(6)
Ql
i Qi
where Qi are the loss Q-factors from N different sources. The horizontal Q-factor is strongly
dependent on the device geometry. In the case of a photonic crystal cavity it is higher for a
larger number of photonic crystal periods. Using an electromagnetic solver [25], we estimated
the horizontal Q-factor by exciting a two-dimensional photonic crystal mode and obtain the
evanescent decay of the mode. We calculated for a 100 × 100 µm2 device a horizontal Q-factor
of approximately 2000 at a wavelength of 8 µm, which can be increased further by making the
resonator larger. Typical waveguide losses of dielectric slabs are in the range of 5 − 20 cm−1 ,
which corresponds to Q-factors in the range of Ql = 103 − 104 . Figure 5 shows the absorption
efficiency as a function of Qc for different values of Ql . If 1/Ql = 0, the critical coupling
condition, changes to 1/Qd = 1/Qc + 1/Ql . Without loss, the maximum absorption efficiency
is independent of Qc . An increasing loss gives a decreasing upper limit for reasonable detector
Q-factors.
To further investigate the consequences of a mismatch between optimally designed Q-factors
and a possible deviation, Fig. 6 shows the absorption efficiency of four devices, each with a
different value of Qc , versus the detector Q-factor Qd . If no loss is included, the absorption
efficiency of each curve has a maximum at Qc = Qd . On a logarithmic x-scale, all peaks have
the same shape and are symmetric, centered around Qc . For a factor of two between optimal
Qd and actual Qd , the absorption efficiency is still at approximately 90 % of its peak value,
independent of Qc . The inclusion of loss reduces the maximum absorption efficiency. However,
the curve retains its shape, thus giving the same impact of a Q-factor mismatch as without
losses.
The maximum performance of a resonant cavity QCD is evaluated for a range of values of
total loss Q-factor in Fig. 7. At zero loss, the responsivity converges to the maximum achievable
responsivity of 107 mA/W, which is obtained from Eq. 1 with η = 0.5, pe = 1 at λ = 8 µm and
30 periods. By making 1/Ql larger, i.e. increasing the total losses, the maximum responsivity is
obtained at lower values of 1/Qd and 1/Qc . When Ql has the same order of magnitude as Qd ,
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6289
Absorption efficiency η
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1/Ql = 0
Qc = 200
(a)
(b)
Ql = 500
Ql = 1000
Ql = 2000
Ql = 4000
Ql = 8000
Qc = 50
Qc = 100
Qc = 200
Qc = 400
10
100
1000
10
100
Detector Q factor Qd
1000
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
Detector Q factor Qd
Fig. 6. Absorption efficiency for different values of Qc versus the QCD Q-factor, without
any losses. The absorption efficiency gets maximal for Qc = Qd . For a mismatch of a factor
of two the absorption efficiency is still at approximately 90 % of its peak value. When loss
is included the limit of the absorption efficiency decreases for higher losses, but the shape
remains, giving the same impact of a Q-factor mismatch as without losses.
Ql = 4000
Ql = 16000
Ql = 8000
Ql = 32000
(a)
2.0x109
(b)
80
1.5x109
60
1.0x109
40
0.5x109
20
0
100
1000
Cavity Q factor Qc
10000
100
1000
Optimum Dj∗ (Jones)
Optimum R (mA/W)
100
Ql = 1000
Ql = 2000
0
10000
Cavity Q factor Qc
Fig. 7. Optimum responsivity and specific detectivity for this QCD design, versus the cavity
Q-factor Qc for different Ql . For each value of Qc , Qd (n2D , d) was optimized for R and
then again for D∗j . For increasing Ql , the responsivity converges against the maximum
achievable responsivity of 107 mA/W for η = 0.5 at this wavelength and 30 periods at T =
300 K. For lower Ql , the optimal performance is only achieved with a QCD of equivalently
lower absorption.
the responsivity decreases significantly. However, the specific detectivity is proportional to the
responsivity and the square root of the resistance. While the responsivity decreases for higher
Ql , the specific detectivity compensates the reduction by an increase of the resistance. The
maximum of the specific detectivity is found at higher values of Qd than for the responsivity.
The most important consequence is, that a reduction of loss, increases the specific detectivity
significantly.
4.
Conclusion and outlook
We have presented a detailed analysis of a resonant cavity quantum cascade photodetector
utilizing a diagonal-transition scheme. The diagonality of the optically active transition was
introduced as new means to obtain quality factor matching between the photodetector and the
resonant cavity. Neglecting losses and for high Q-factors, the maximum of the responsivity
converges to its highest possible value for η = 0.5 and the chosen number of periods. The
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6290
maximum of the specific detectivity gets arbitrarily large for increasing Q-factors. In reality
such high values are inhibited by losses. We calculated the influence of varying losses on the
device performance to obtain reasonable values for a detector Q-factor to design QCDs. The
impact of a mismatch between designed Q-factors and possible deviations was analyzed. A
deviation by a factor of two of Qc from its optimum value for given Ql and Qd , but keeping the
correct resonance position, reduces the absorption efficiency by only 10%. The responsivity and
specific detectivity was calculated for a range of loss values. The calculation predicts a possible
responsivity of 95 mA/W and a specific detectivity of approximately 109 Jones at T = 300 K
for QCDs with a diagonal optically active transition and optimized Q-factors. The optimum
detectivity was obtained with a sheet carrier density of n2D = 4.9x1010 cm−2 and a barrier
width of d = 4.8 nm. The resonance width is approximately 0.3 cm−1 .
Acknowledgments
The authors acknowledge the support by the Austrian Science Funds (FWF) in the framework
of the Doctoral School Building Solids for Function (project W1243), the project NextLite
(SFB F49-09) and the FP7 EU-project ICARUS.
#229011 - $15.00 USD
(C) 2015 OSA
Received 4 Dec 2014; revised 28 Jan 2015; accepted 28 Jan 2015; published 2 Mar 2015
9 Mar 2015 | Vol. 23, No. 5 | DOI:10.1364/OE.23.006283 | OPTICS EXPRESS 6291