You are on page 1of 7

Journal of Luminescence 239 (2021) 118393

Contents lists available at ScienceDirect

Journal of Luminescence
journal homepage: www.elsevier.com/locate/jlumin

Influence of the doping type on the temperature dependencies of the


photoluminescence efficiency of InGaAlAs/InGaAs/InP heterostructures
Innokenty I. Novikov a, *, Alexey M. Nadtochiy b, Aleksandr Yu Potapov c, Andrey G. Gladyshev a,
Evgenii S. Kolodeznyi a, Stanislav S. Rochas a, Andrey V. Babichev a, Vladislav V. Andryushkin a,
Dmitriy V. Denisov d, Leonid Ya Karachinsky a, Anton Yu Egorov e, f, Vladislav E. Bougrov a
a
ITMO University, St. Petersburg, 197101, Russia
b
National Research University Higher School of Economics, St. Petersburg, 190008, Russia
c
Peter the Great St.Petersburg Polytechnic University, St. Petersburg, 195251, Russia
d
Saint Petersburg Electrotechnical University “LETI”, St. Petersburg, 197376, Russia
e
Connector Optics LLC, St. Petersburg, 194292, Russia
f
Alferov University, St. Petersburg, 194021, Russia

A R T I C L E I N F O A B S T R A C T

Keywords: The temperature dependencies of photoluminescence efficiency of InGaAlAs/InGaAs/InP heterostructures with


Molecular beam epitaxy 1550 nm range strained quantum wells have been studied. The heterostructures had nine In0.74Ga0.26As quantum
Quantum well wells which were separated by In0.53Al0.20Ga0.27As barriers. The barriers were δ-doped with n- and p-type
Photoluminescence
dopants with different layer concentrations of charge carriers. The both p-type and n-type doping of barriers
Heterostructures
leads to slight increase in the efficiency when doping level not exceeding values of (1–5)⋅1012 cm− 2. Moreover,
doping with an n-type dopant leads to a significant increase in the width of the photoluminescence spectrum,
while doping with a p-type dopant leads to a decrease in its width, compared to a heterostructure with undoped
barrier layers.

1. Introduction transparence carrier density and improve the differential gain [8], as
well as reduce the broadening inhomogeneity. At the same time, it was
Lasers of 1550 nm spectral range are of great interest due to the shown that along with this, doping leads to deterioration in modulation
presence of a minimum of optical losses in the optical fiber, as well as the characteristics, limiting the performance at the level of 10 Gb/s [9].
presence of effective erbium fiber amplifiers [1] which provide ampli­ Recently it has been shown that an increase of acceptors concentration
fication of optical signals without distortion. Significant advances have leads to an increase in gain compression due to an increase of internal
been achieved in the creation of lasers of this spectral range on quantum losses in the active region and thus limits the maximum modulation
dots [2]. However, structures based on InGaAsP/InP materials, which bandwidth [10] Other limiting factors for p-doped QD-based lasers
are produced by metal organic chemical vapor deposition (MOCVD), are include a decrease in modal enhancement due to the inter-valance band
usually used to create lasers operating at wavelengths of 1520–1580 nm absorption and increased nonradiative Auger recombination [11–13].
[3]. The disadvantage of this material system is the low energy of Using p-doping, GaAs-based substrate QD-based lasers operating at
electron localization in quantum wells (QWs). The concept of doping 1100 and 1300 nm have achieved a maximum 3-dB bandwidth of 11 and
with respect to QWs structures is well known and was proposed in Refs. 8 GHz, respectively, which are only higher than those of the undoped
[4,5]. In particular, it was assumed that the development of this devices [14].
approach would be most applicable to quantum dots (QD)-based lasers For the case of QWs active regions, the use of n-type modulation
due to the lower density of states in them [6]. By n-type doping the doped (MD) multiple QWs (MQWs) structures (with doping of barrier
increase of linewidth enhancement factor 1300 nm QD-based lasers is layers) allows to lower the threshold current density as for the case of
realized [7]. Using a p-doping in QD-based lasers would reduce MOCVD-grown 1300-nm [15], 1500-nm InGaAsP/InP-based lasers [16],

* Corresponding author.
E-mail address: novikov@switch.ioffe.ru (I.I. Novikov).

https://doi.org/10.1016/j.jlumin.2021.118393
Received 29 March 2021; Received in revised form 16 July 2021; Accepted 10 August 2021
Available online 11 August 2021
0022-2313/© 2021 Elsevier B.V. All rights reserved.
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

and for MBE-grown 1300-nm InAsP/InP-based lasers [17–20]. At the a barrier for charge carriers. The In0.53Ga0.27Al0.20As layer with the band
same time, by p-type doping the enhancement of the differential gain, T0 gap ~1.03 eV was used as the confinement and barrier layers. The active
characteristic temperature and the reduction of the turn-on delay time region was grown between the In0.53Ga0.27Al0.20As confinement layers
for MOCVD-grown 1500-nm GaInAs/InP-based [5], 1300-nm and consisted of nine 3.7-nm-thick In0.74Ga0.26As QWs. The QWs were
InGaAsP/InP-based [21] as well as for MBE-grown 1300-nm separated by 12-nm-thick δ-doped In0.53Ga0.27Al0.20As barriers (see
InAsP/InP-based lasers [18,20] are observed. Consequently, p-type Fig. 1). The heterostructures differed in the value and type of doping of
MD MQWs active region is promising from the point of view of imple­ the barrier layers. Structure “U19” was grown as a reference sample with
mentation high-speed lasers. Moreover, the reduction of the threshold undoped barriers. The “Si10” and “Si50” samples were grown with
carrier densities in the n-type MD-MQW laser can be explained by the silicon-doped barrier layers with a doping concentration of 1⋅1012cm− 2
reduction of intervalence band absorption loss. In opposite, the p-type and 5⋅1012cm− 2 respectively. The “C10”, “C20” and “C50” samples were
doping increase the internal loss due to increase in the absorption loss grown with carbon-doped barrier layers with a doping concentration of
caused by the increase of the intervalence band absorption [18]. 1⋅1012cm− 2, 2⋅1012cm− 2 and 5⋅1012cm− 2 respectively. Carbon was used
Beryllium (Be) in comparison with zinc (Zn) has a lower diffusion as a p-type dopant because it does not diffuse, unlike the commonly used
coefficient, and, as a result, allows to localize impurity atoms to a greater beryllium. δ-Doping was chosen due to reproducibility and high accu­
extent in barrier layers in comparison with Zn, which allows to reduce racy of the concentrations of the dopants. The heterostructures were
excess absorption in the active layer [20]. Moreover, usage of carbon (C) grown using two gallium, two aluminum and one indium sources.
instead of Be allows additional localization of impurities in barriers [22, Growth rates of gallium sources were 0.5 and 0.35 Å/s, aluminum
23]. sources – 0.38 and 0.92 Å/s, indium source – 1 Å/s. The growth tem­
At this time, by application of p-doped barriers the increase of perature during the growth process was 630 ◦ C. The description of
vertical-cavity surface-emitting laser (VCSELs) modulation bandwidth is grown heterostructures is summarized in Table 1.
realized. The 12 Gbps VCSELs modulation speed is realized [24]. The active region was covered on the surface side by a 30-nm-thick
Moreover, it was realized the increase of -3dB bandwidth to 21 GHz in In0.52Al0.48As layer, which was covered by a thin In0.53Ga0.47As cap
relation with undoped case of 850 nm VCSELs (18 GHz) [25]. Here, we layer to prevent surface oxidation. Fig. 1 shows the schematic repre­
present the results of complex studies of the optical properties of sentation of the grown structures and band diagram.
InGaAlAs/InGaAs/InP heterostructures with strained InGaAs QWs. The The excitation of photoluminescence (PL) was performed by a
barrier layers of the heterostructures were selectively doped (carbon or YAG: Nd laser with a wavelength of 527 nm and maximum output power
silicon) to the values of layer concentrations (1–5)⋅1012 cm− 2. The re­ of 500 mW. The optical power of the laser was adjusted using neutral
sults of this study can be applied for the improve the performance light filters. The diameter of the focused laser spot is ~30 μm. The
long-wavelength VCSELs [26,27]. measurements were carried out in a cryostat with temperature

2. Experiment
Table 1
For our experiments, six semiconductor heterostructures with The description of the grown heterostructures.
strained InGaAs QWs were grown on semi-insulated InP (100) substrates Structure C10 C20 C50 U19 Si10 Si50
using a Riber 49 molecular beam epitaxy (MBE) system. A 200-nm Dopant С С С – Si Si
Concentration, 1012 cm− 2
1.0 2.0 5.0 0.0 1.0 5.0
In0.52Al0.48As buffer layer was grown on the InP substrate and used as

Fig. 1. The schematic representation of the grown structures and band diagram.

2
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

stabilization in the range of 13–325 K. A germanium photodiode cooled dependence where radiative and nonradiative recombination rates are
by liquid nitrogen is used as a detector. The lock-in technique was used equal (Bn2 = An). The lower excitation level corresponding to the in­
for characterization. flection point indicates the lower the excitation levels where the radi­
ative recombination (Bn2) begins to dominate, and therefore the optical
3. Results and discussion quality of the structure is higher.
Undoped sample “U19” was approximated by equation (1) and the
Fig. 2 displays the PL spectra of the six heterostructures at room inflection point has been defined (see Fig. 3). PL of this sample dem­
temperature. The excitation power density is 600 W/cm2. onstrates superlinear growth at low excitation densities (nonradiative
It can be seen that PL peak position corresponding to the InGaAs QWs recombination dominates), which transits to linear growth (radiative
for all samples has the same energy of 0.810 eV (1530 nm). This cor­ recombination dominates) of excitation densities above 1850 W/cm2
responds to the transition energy from the ground state level of electrons (see Table 2). However, the doped samples demonstrate a qualitatively
(e1) to the ground state level of heavy holes (hh1) in QWs. There is also a different behavior of the PL intensity with an increase of the excitation
slight inflection of the high-energy part of the “Si 50” structure spec­ density: the dependencies do not have a pronounced superlinear growth
trum, which occurs due to the radiative recombination in the barrier area. To describe this behavior, it is necessary to supplement the model
layers. Fig. 2 shows that the samples have different peak and integrated with the mechanism of recombination of nonequilibrium charge carriers
PL intensities. The highest intensity at the peak of the PL spectrum is with dopant. Nonequilibrium charge carriers are born as a result of
shown by the “C10” sample doped with carbon to the level of 1.0 * optical excitation. Doping charge carriers are a result of the ionization of
1012cm− 2, although the difference is rather small. dopant impurities. In this case, the radiative recombination of charge
The standard equation for the charge carrier recombination rate (R) carriers (Bn2) must be supplemented by the linear term n0n, where n0 is
is used for the analysis of recombination processes [28]. It is assumed the coefficient of the concentration of dopant impurities [29]. Equation
that the system is in equilibrium and the recombination rate is equal to (1) for the approximation of doped samples can be written as:
the rate of charge carrier generation. In this case, the generation rate is
Pex ∼ An + n0 n + Bn2 , (2)
proportional to the excitation power density Pex:

Pex ∼ R = An + Bn2 + Cn3 , (1) where sum (n0n + Bn2) is radiative recombination and we neglected
Auger mechanism. As a result of the approximation by equation (2), two
Where n is the charge carrier concentration (the concentration of inflection points on the dependence can be distinguished. The first in­
electrons is equal to the concentration of holes), An is Shockley-Read- flection point at which the radiative recombination (n0n + Bn2) is equal
Hall (SRH) recombination rate, Bn2 is the radiative recombination to the non-radiative recombination (An). The second inflection point on
rate, Cn3 is the rate of Auger recombination. According to equation (1), the dependence indicated where linear (n0n) and quadratic (Bn2) com­
at low levels of excitation, the SRH recombination dominates, and at ponents of the radiative recombination are equal.
high levels, the Auger mechanism dominates. At intermediate levels of In this paper, we were able to approximate the dependencies of all
excitation, the radiative recombination process can be expected to be doped structures using Equation (2) (see Fig. 3). Samples “C10”, “C20”,
dominant. The contribution of various recombination processes can be and “Si10” have the highest integral PL intensity over the entire range of
calculated by approximating the dependence of the integral PL intensity excitation power densities. The undoped structure has an almost one-
on the excitation power density (see Fig. 3) by equation (1). The order-of-magnitude lower integral PL intensity compared to samples
contribution of Auger recombination at these excitation levels is insig­ “C10”, “C20”, and “Si20” at low excitation densities. However, an
nificant. The approximation will be more clearly if we change the co­ equalization of the integral PL intensities is observed at a high level of
ordinate axes to the laser power density depends vs integral PL. As a excitation.
result of the approximation, we can determine the inflection point of Table 2 shows the obtained values of the excitation power density at

Fig. 2. PL spectra of samples measured at 300 K. The excitation power density is 600 W/cm2. The PL intensity scale is represented on a semi-logarithmic scale.

3
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

Fig. 3. Dependence of the excitation power on the integral PL intensity and corresponding approximation curves. The curves are shifted along the horizontal axis for
better visibility. Solid symbols show the inflection points of the dependencies calculated as results of the approximation.

the inflection points of the dependencies (see Fig. 3). In heterostructures


Table 2
with doped barrier layers, as expected we see the increase of value of
The excitation power density in the inflection points of the dependencies
power density where transition from linear (impurity) optical recom­
(Fig. 3.).
bination to quadratic optical recombination is occurred, as well as we
Structure U19 C10 C20 C50 Si10 Si50 observe a decrease of excitation power density at which radiative
Pex at An = n0n + Bn2, W/cm2 1850 66 60 400 78 28 recombination begins to dominate over nonradiative. At doping level of
Pex at n0n = Bn2, W/cm2 – 11 12 87 63 2 1–2∙1012 cm− 2, the radiative recombination dominates at 60–80 W/
cm2, and these structures exhibit bright PL over the entire range of
excitation densities. A further increase in doping leads to a noticeable
quenching of luminescence and a corresponding increase in the power

Fig. 4. Temperature dependencies of integral PL intensity at the excitation power density of 600 W/cm2, on a semi-logarithmic scale.

4
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

density of the radiative recombination dominance (except for the (see Fig. 6a). As the temperature decreases, an extra peak (B) appears in
structure “Si50”). the high-energy part of the PL spectrum. Fig. 6b shows that at low
Thus, with moderate doping levels, the efficiency of photo­ temperature, the spectra are reproduced in a wide range of the excita­
luminescence increases at low excitation levels, which may be con­ tion power density and do not change its shape. Also, the radiation in­
nected to a change in time of carrier capture into QWs. tensity of the barrier layer in relation to the radiation intensity of the
Fig. 4 shows the temperature dependencies of integral PL intensity at QWs (A) does not change with the excitation power density. It means
the excitation power density of 600 W/cm2. that the probabilistic recombination processes in the QWs and barrier
When the temperature increases from 13 to 325 K, the drop of the are independent on the pump level. The appearance of PL from barrier
integral PL intensity is different for all samples. The sample with the layers is not related to the saturation of the PL from QWs, since in this
highest temperature stability among the studied structures is the “C10” case, the PL from barrier layers should be attenuated with a decrease in
sample which is doped by carbon with concentration of 1⋅1012 cm− 2. It excitation. It is assumed that PL from barrier layers is associated with the
can also be seen that the “C20” sample with doping level twice as high as bending of the energy zones near the barrier-well interface due to high
“C10” has almost identical stability. To obtain the temperature depen­ doping level and the deterioration of the charge carriers transport with a
dence of the width of the PL spectrum, the integral characteristic of the decrease in temperature. As a result, charge carriers are accumulated in
spectrum (Δhν) was used. This characteristic is the root mean square the barrier layers and then recombine with emitting of quanta whose
deviation of the quantum energy (hν) from the mathematical expecta­ energy is equal to the energy of the band gap of the barrier layers.
tion (RMS). The integral characteristic of the spectrum is used in the
study of complex, noisy or asymmetrical waveforms. This approach al­ 4. Conclusion
lows us to obtain a more stable estimation of the PL spectrum width,
compared to the using of full width at half maximum (FWHM). The root In summary, the temperature dependencies of photoluminescence
mean square deviation Δhν was calculated by equation (3): efficiency of InGaAlAs/InGaAs/InP heterostructures with 1550 nm
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ range strained quantum wells have been studied. The analysis of the
√∫ +∞
√ PL(hν)⋅(hνmean − hν)2 dhν integral PL intensities for all studied heterostructures allowed to define
Δhν = RMS = √ − ∞ ∫ +∞ , (3)
PL(hν)dhν the values of the inflection points and the corresponding excitation
− ∞
power densities where the radiative and non-radiative recombination
where PL(hν) is the PL intensity for the quantum energy hν, hνmean is the are equal. It was found that the moderate doping (up to cm− 2) of the
mathematical expectation of the quantum energy hν. The hvmean is barrier layers increases the efficiency of radiative recombination at low
calculated using equation (4): excitation levels. The dependencies of the spectrum width on the exci­
∫ +∞ tation power density were analyzed. The width of the spectra of the
PL(hν)⋅hνdhν doped heterostructures exceeds the width of the spectrum of the undo­
hνmean = −∫∞+∞ . (4)
PL(hν)dhν ped sample. This fact is related to the levels of holes that are occupied by
− ∞
charge carriers from doped barrier layers. The appearance of the second
Fig. 5 shows the temperature dependencies of the RMS of the PL peak at PL spectrum with a decrease in temperature was found for
spectra after processing the experimental data. samples doped with silicon up to 5⋅1012 cm− 2. The second peak is
Fig. 6 shows PL spectra of the “Si50” sample at different tempera­ associated with the radiation of the InGaAlAs confinement or barrier
tures and PL spectra of the “Si50” sample at 13 K with different exci­ layers. It is assumed that this fact is associated with the bending of the
tation levels. energy zones near the barrier-quantum well interface due to the high
The samples “Si10” and “Si50” have the highest RMS values. This concentration of silicon doping. Obtained results can be used in the
fact is associated with the recombination of carriers in the barrier layer development of light-emitting structures for 1550 nm spectral range.

Fig. 5. Temperature dependencies of the RMS of the PL spectra. Excitation power density 600 W/cm2.

5
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

Fig. 6. (a) PL spectra of the “Si50” sample at different temperatures. (b) PL spectra of the “Si50” sample at 13 K and different excitation levels.

Silicon doping may be promising for creating lasers and super­ Conceptualization and Supervision. Vladislav E. Bougrov: Supervision,
luminescent diodes tunable in a wide spectral range. Project administration and Writing - Review & Editing.

Author statement Declaration of competing interest

Innokenty I. Novikov: Supervision and Writing - Original Draft. The authors declare that they have no known competing financial
Alexey M. Nadtochiy: Investigation, Formal analysis, Data curation and interests or personal relationships that could have appeared to influence
Funding acquisition. Aleksandr Yu. Potapov: Investigation. Andrey G. the work reported in this paper.
Gladyshev: Methodology. Evgenii S. Kolodeznyi: Resources. Stanislav S.
Rochas: Resources. Andrey V. Babichev: Formal analysis and Writing - Acknowledgments
Original Draft. Vladislav V. Andryushkin: Writing – original draft and
Investigation. Dmitriy V. Denisov: Methodology. Leonid Ya. Kar­ The work was carried out with the support of the Ministry of Science
achinsky: Writing – review & editing and Validation. Anton Yu. Egorov: and Higher Education of the Russian Federation (research project No.

6
I.I. Novikov et al. Journal of Luminescence 239 (2021) 118393

2019-1442). N. A. M. thanks for the support of the Program of Funda­ [15] K. Nakahara, K. Uomi, T. Tsuchiya, A. Niwa, T. Haga, T. Taniwatari, 1.3-/spl mu/m
InGaAsP-InP n-type modulation-doped strained multiquantum-well lasers, IEEE J.
mental Research of the National Research University "Higher School of
Sel. Top. Quant. Electron. 3 (1997) 166–172, https://doi.org/10.1109/
Economics" in terms of measurements. 2944.605650.
[16] T. Yamamoto, T. Watanabe, S. Ide, I. Tanaka, H. Nobuhara, K. Wakao, Low
References threshold current density 1.3-μm strained-layer quantum-well lasers using n-type
modulation doping, IEEE Photon. Technol. Lett. 6 (1994) 1165–1166, https://doi.
org/10.1109/68.329626.
[1] A.W. Naji, B.A. Hamida, X.S. Cheng, M.A. Mahdi, S. Harun, S. Khan, W.F. Al- [17] H. Shimizu, K. Kumada, N. Yamanaka, N. Iwai, T. Mukaihara, A. Kasukawa, Low
Khateeb, A.A. Zaidan, B.B. Zaidan, H. Ahmad, Review of Erbium-doped fiber threshold 1.3 [micro sign] m InAsP n-type modulation doped MQW lasers grown
amplifier, Int. J. Phys. Sci. 6 (2011) 4674–4689, https://doi.org/10.5897/ by gas-source molecular-beam epitaxy, Electron. Lett. 34 (1998) 888, https://doi.
IJPS11.782. org/10.1049/el:19980662.
[2] I.I. Novikov, N.Yu Gordeev, M.V. Maximov, YuM. Shernyakov, A.E. Zhukov, A. [18] H. Shimizu, K. Kumada, N. Yamanaka, N. Iwai, T. Mukaihara, A. Kasukawa, 1.3-/
P. Vasil’ev, E.S. Semenova, V.M. Ustinov, N.N. Ledentsov, D. Bimberg, N. spl mu/m InAsP modulation-doped MQW lasers, IEEE J. Quant. Electron. 36
D. Zakharov, P. Werner, Ultrahigh gain and non-radiative recombination channels (2000) 728–735, https://doi.org/10.1109/3.845730.
in 1.5 μm range metamorphic InAs–InGaAs quantum dot lasers on GaAs substrates, [19] H. Shimizu, K. Kumada, N. Yamanaka, N. Iwai, T. Mukaihara, A. Kasukawa,
Semicond. Sci. Technol. 20 (1) (2004) 33, https://doi.org/10.1088/0268-1242/ Submilliampere threshold current in 1.3/spl mu/m InAsP n-type modulation doped
20/1/005. MQW lasers grown by gas source molecular beam epitaxy, Electron. Lett. 34 (1998)
[3] J. Zhu, MOCVD Growth of III-V Compounds for Long Wavelength Optoelectronic 1591–1593, https://doi.org/10.1049/el:19981076.
Devices, Master’s Thesis, Nanyang Technological University, Singapore, 2005, [20] H. Shimizu, K. Kumada, N. Yamanaka, N. Iwai, T. Mukaihara, A. Kasukawa, 1.3 μm
https://doi.org/10.32657/10356/4075. InAsP p-type modulation doped MQW lasers grown by gas-source molecular-beam
[4] Kerry J. Vahala, C.E. Zah, Effect of doping on the optical gain and the spontaneous epitaxy, J. Cryst. Growth 201 (1999) 896–900, https://doi.org/10.1016/s0022-
noise enhancement factor in quantum well amplifiers and lasers studied by simple 0248(98)01488-2.
analytical expressions, Appl. Phys. Lett. 52 (23) (1988) 1945–1947, https://doi. [21] F. Kano, T. Yamanaka, N. Yamamoto, Y. Yoshikuni, H. Mawatari, Y. Tohmori,
org/10.1063/1.99584. M. Yamamoto, K. Yokoyama, Reduction of linewidth enhancement factor in
[5] C.E. Zah, R. Bhat, S.G. Menocal, F. Favire, N.C. Andreadakis, M.A. Koza, C. Caneau, InGaAsP-InP modulation-doped strained multiple-quantum-well lasers, IEEE J.
S.A. Schwarz, Y. Lo, T.P. Lee, Cavity length and doping dependence of 1.5-mu m Quant. Electron. 29 (1993) 1553–1559, https://doi.org/10.1109/
GaInAs/GaInAsP multiple quantum well laser characteristics, IEEE Photon. islc.1992.763558.
Technol. Lett. 2 (4) (1990) 231–233, https://doi.org/10.1109/68.53245. [22] S. Godey, S. Dhellemmes, A. Wilk, M. Zaknoune, F. Mollot, CBr4 and Be heavily
[6] N.T. Yeh, J.M. Lee, T.E. Nee, J.I. Chyi, Self-assembled In/sub 0.5/Ga/sub 0.5/As doped InGaAs grown in a production MBE system, J. Cryst. Growth 278 (2005)
quantum-dot lasers with doped active region, IEEE Photon. Technol. Lett. 12 (9) 600–603, https://doi.org/10.1016/j.jcrysgro.2004.12.075.
(2000) 1123–1125, https://doi.org/10.1109/68.874209. [23] P.H. Lei, C.D. Yang, M.Y. Wu, M.C. Wu, K.Y. Cheng, C.C. Lin, W.J. Ho, Effects of n-
[7] Y.Q. Qiu, Z.R. Lv, H. Wang, H.M. Wang, X.G. Yang, T. Yang, Improved linewidth type modulation-doping barriers and a linear graded-composition GaInAsP
enhancement factor of 1.3-μ m InAs/GaAs quantum dot lasers by direct Si doping, intermediate layer on the 1.3 μ m AlGaInAs∕ AlGaInAs strain-compensated
AIP Adv. 11 (2021), 055002, https://doi.org/10.1063/5.0044313. multiple-quantum-well laser diodes, J. Vac. Sci. Technol. B: Microelectron.
[8] Z. Zhang, D. Jung, J.C. Norman, P. Patel, W.W. Chow, J.E. Bowers, Effects of Nanometer. Struct. Process. Measure Phenomena 24 (2006) 623–628, https://doi.
modulation p doping in InAs quantum dot lasers on silicon, Appl. Phys. Lett. 113 org/10.1116/1.2172954.
(2018), 061105, https://doi.org/10.1063/1.5040792. [24] N. Hatori, A. Mizutani, N. Nishiyama, A. Matsutani, T. Sakaguchi, F. Motomura,
[9] K. Otsubo, N. Hatori, M. Ishida, S. Okumura, T. Akiyama, Y. Nakata, H. Ebe, F. Koyama, K. Iga, An over 10-Gb/s transmission experiment using a p-type delta-
M. Sugawara, Y. Arakawa, Temperature-insensitive eye-opening under 10-Gb/s doped InGaAs-GaAs quantum-well vertical-cavity surface-emitting laser, IEEE
modulation of 1.3-μm p-doped quantum-dot lasers without current adjustments, Photon. Technol. Lett. 10 (1998) 194–196, https://doi.org/10.1109/68.655355.
Jpn. J. Appl. Phys. 43 (8B) (2004), https://doi.org/10.1143/JJAP.43.L1124. [25] K.L. Chi, D.H. Hsieh, J.L. Yen, X.N. Chen, J. Chen, H.C. Kuo, Y.J. Yang, J.W. Shi,
L1124. 850 nm VCSELs with P-type δ-doping in the active layers for improved high-speed
[10] A. Martinez, Y. Li, L.F. Lester, A.L. Gray, Microwave frequency characterization of and high-temperature performance, IEEE J. Quant. Electron. 52 (2016) 1–7,
undoped and p-doped quantum dot lasers, Appl. Phys. Lett. 90 (25) (2007) 251101, https://doi.org/10.1109/JQE.2016.2611439.
https://doi.org/10.1063/1.2749432. [26] A.V. Babichev, et al., 6-mW single-mode high-speed 1550-nm wafer-fused VCSELs
[11] M.L. Huberman, A. Ksendzov, A. Larsson, R. Terhune, J. Maserjian, Optical for DWDM application, IEEE J. Quant. Electron. 53 (2017) 1–8, https://doi.org/
absorption by free holes in heavily doped GaAs, Phys. Rev. B 44 (3) (1991) 1128, 10.1109/JQE.2017.2752700.
https://doi.org/10.1103/PhysRevB.44.1128. [27] S. Blokhin, A. Babichev, A. Gladyshev, L. Karachinsky, I. Novikov, A. Blokhin,
[12] I.P. Marko, N.F. Massé, S.J. Sweeney, A.D. Andreev, A.R. Adams, N. Hatori, S. Rochas, D. Denisov, K. Voropaev, A. Ionov, N. Ledentsov, A. Egorov, Wafer-
M. Sugawara, Carrier transport and recombination in p-doped and intrinsic 1.3 μ m fused, 1300 nm VCSELs with an active region based on superlattice, Electron. Lett.
in As∕ Ga as quantum-dot lasers, Appl. Phys. Lett. 87 (21) (2005) 211114, https:// (2021), https://doi.org/10.1049/ell2.12232.
doi.org/10.1063/1.2135204. [28] A.M. Nadtochiy, M. V Maximov, S.A. Mintairov, N.A. Kalyuzhnyy, S. Rouvimov, A.
[13] I.P. Marko, A.R. Adams, S.J. Sweeney, D.J. Mowbray, M.S. Skolnick, H.Y. Liu, K. E. Zhukov, InGaAs/GaAs hybrid quantum well-dot nanostructures: impact of
M. Groom, Recombination and loss mechanisms in low-threshold InAs-GaAs 1.3-/ substrate orientation and recombination mechanisms, J. Phys. Conf. 917 (2017),
spl mu/m quantum-dot lasers, IEEE J. Sel. Top. Quant. Electron. 11 (5) (2005) 032001, https://doi.org/10.1088/1742-6596/917/3/032001.
1041–1047, https://doi.org/10.1109/JSTQE.2005.853847. [29] R. Olshansky, C. Su, J. Manning, W. Powazinik, Measurement of radiative and
[14] S. Fathpour, Z. Mi, P. Bhattacharya, Small-signal modulation characteristics of p- nonradiative recombination rates in InGaAsP and AlGaAs light sources, IEEE J.
doped 1.1-and 1.3-μm quantum-dot lasers, IEEE Photon. Technol. Lett. 17 (11) Quant. Electron. 20 (1984) 838–854, https://doi.org/10.1109/
(2005) 2250–2252, https://doi.org/10.1109/LPT.2005.857242. JQE.1984.1072500.

You might also like