Вы находитесь на странице: 1из 5

Journal of Luminescence 155 (2014) 2731

Contents lists available at ScienceDirect

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

Shell to core carrier-transfer in MBE-grown GaAs/AlGaAs coreshell


nanowires on Si(1 0 0) substrates
Maria Herminia Balgos n, Rafael Jaculbia, Michael Defensor, Jessica Pauline Afalla, Jasher
John Ibaes, Michelle Bailon-Somintac, Elmer Estacio, Arnel Salvador, Armando Somintac
Condensed Matter Physics Laboratory, National Institute of Physics, University of the Philippines, Diliman, Quezon City 1101, Philippines

art ic l e i nf o a b s t r a c t

Article history: We report on the shell-to-core carrier-transfer in GaAs/Al0.1Ga0.9As core-shell nanowires grown on
Received 26 August 2013 Si(1 0 0) substrates via molecular beam epitaxy. The nanowires are dominantly zincblende and are tilted
Received in revised form with respect to the substrate surface. Photoluminescence (PL) excitation spectrosocopy at 77 K revealed an
2 June 2014
abrupt increase in the GaAs PL intensity at excitation above the Al0.1Ga0.9As shell bandgap which is attributed
Accepted 5 June 2014
Available online 16 June 2014
to shell to core carrier-transfer. More carriers from the Al0.1Ga0.9As transfer to the GaAs at T490 K, as
observed in the time-resolved PL and temperature dependence of the relative PL intensities of GaAs and
Keywords: Al0.1Ga0.9As due to the ionization of the traps within the Al0.1Ga0.9As. Using a coupled rate equation model
IIIV semiconductors that takes into account shell to core carrier-transfer, the average recombination time constants of Al0.1Ga0.9As
Carrier transfer
shell rec,s 400 ps (580 ps) and GaAs core rec,c 600 ps (970 ps) were obtained from the time-resolved PL at
Nanowires
300 K (77 K). Carrier-transfer time constants CT 50 ps (55 ps) at 300 K (77 K) were also obtained.
Time-resolved luminescence
& 2014 Elsevier B.V. All rights reserved.

1. Introduction The enhancement of the PL intensity in CSNWs is attributed


mainly to the reduced interaction of carriers with the surface
Semiconductor nanowires (NWs) are quasi-one dimensional states which may act as nonradiative recombination sites, electro-
structures with diameters in nanometers and length in microns. nic traps, and scattering sites [11,12,19]. However, the introduction
The wide range of base materials and the possible production of of the shell can also enhance the PL intensity by acting as another
highly crystalline structures from lattice-mismatched materials source of carriers for the NW core [20]. The carriers from the shell
make NWs a versatile alternative to its bulk counterpart [1-3]. In can transfer to the core and effectively increase the number of
fact, the integration of IIIV semiconductors to Si has been possible recombining carriers thereby producing a more intense PL [21].
through NW growth [47]. Due to its interesting properties, NWs Carrier transfer (CT) has already been observed in quantum wells
cover a wide range of applications such as in energy conversion, [2223], laser structures [24], and type II coreshell heterostruc-
sensors, electronics, and optoelectronics. tures [20] using time resolved PL (TRPL).
Application of NWs as light emitters such as light emitting In this work, we study in detail the shell-to-core CT in GaAs/
diodes [8,9] and lasers [10] has been previously demonstrated. Due Al0.1Ga0.9As CSNWs by utilizing continuous wave (CW) and TRPL
to their high surface-to-volume ratio, however, NWs are suscep- for varying temperatures. The relatively low Al mole fraction
tible to surface states that may degrade their light emission (x 0.1) of the AlxGa1  xAs allows us to observe a distinct PL signal
capabilities [11,12]. To reduce the effects of surface states, NWs from GaAs and AlGaAs and at the same time, to pump both GaAs
are coated with high bandgap material (shell) that serves as a and AlGaAs using a Ti:Sapphire CW laser. The effect of the CT on
passivation barrier from the surface states. Electrical and optical the PL intensity is measured directly using PL excitation (PLE).
investigations in these coaxial structures, commonly known as
core-shell NWs (CSNWs), show enhanced carrier mobility [13,14]
and photoluminescence (PL) intensity [1518], as compared to 2. Methods
bare NWs (no shell).
GaAs/Al0.1Ga0.9As CSNWs coated with n-doped GaAs cap layer
were grown on Si (1 0 0) substrates using a Riber 32P molecular
beam epitaxy (MBE) machine. The Si substrate was pre-patterned
n
Corresponding author. Tel./fax: 63 99209749. with gold nanodots that serve as catalysts for the vaporliquidsolid
E-mail address: mbalgos@nip.upd.edu.ph (M.H. Balgos). (VLS) growth of NWs [25]. GaAs was deposited at 580 1C for 30 min

http://dx.doi.org/10.1016/j.jlumin.2014.06.008
0022-2313/& 2014 Elsevier B.V. All rights reserved.
28 M.H. Balgos et al. / Journal of Luminescence 155 (2014) 2731

followed by Al0.3Ga0.7As at the same growth temperature and


duration. The thin lm equivalent thickness is 0.65 m and
0.84 m for GaAs and Al0.3Ga0.7As, respectively. Lastly, a 0.1 m Si-
doped GaAs cap ( 1017 cm  3) served as a protective coating for the
AlGaAs which readily oxidizes when exposed to air [26]. The n-
doping of
the GaAs cap decreases the width of the depletion region at the
airAlGaAs interface [13,14] preventing it from penetrating to
the AlGaAs. The formation of a depletion shell could narrow the
electronic channel producing PL [15] and decreases the carrier
lifetime [5].
CW-PL measurements were conducted at temperatures of
10300 K. The sample was mounted on the coldnger of a
closed-cycle He cryostat equipped with a Lakeshore temperature
controller. The excitation source was a 488 nm Ar laser and the
PL signal was dispersed by a SPEX 500 m monochromator and Fig. 2. 300 K PL of GaAs/AlGaAs core-shell nanowires and bare GaAs nanowires.
The (  10) means that the PL spectrum of GaAs is multiplied by 10 for clarity. More
detected by a Hamamatsu GaAs photomultiplier tube. PLE (at 77 K)
than an order of magnitude increase in the GaAs PL peak intensity of the coreshell
and TRPL (at 77 K and 300 K) measurements were conducted at a nanowires compared to bare nanowires is observed. The inset is a schematic of the
separate -PL setup where a 10  objective was used to focus the coreshellcap structure.
laser and collect the PL signal from the sample mounted inside a
continuous ow liquid nitrogen cryostat for 77 K measurements.
For the PLE, the GaAs PL peak intensity at 1.514 eV was monitored 141 714 nm with a density of 1.68  108 cm  2. The GaAs core
using a Synapse Si CCD detector attached to an iHR550 Horiba diameter and the AlGaAs shell thickness is  100 nm and  19 nm,
Jobin Yvon spectrometer. The excitation energy of a CW Ti: respectively. These values were estimated using the diameter of a
Sapphire laser was tuned from 1.531.68 eV (810740 nm) corre- bare GaAs NW (1037 25 nm) grown separately using the same
sponding to below and above the Al0.1Ga0.9As shell bandgap, parameters as reference. For MBE-grown NWs via the VLS
respectively. For the TRPL measurements, a frequency doubled mechanism, the growth elements are decomposed as they are
Ti:Sapphire femtosecond laser (400 nm) with 100 fs pulses at deposited on the substrate. Hence, a lm is also formed in areas
80 MHz repetition rate was used as excitation. The uence was with no gold catalysts [30]. However, due to the high lattice
kept at 0.012 J/cm3 to ensure minimal, if any, nonlinear effects mismatch between GaAs and Si (4%) [31], this lm is defective
[27]. The PL signal was fed to a streak camera system consisting of and emits very weak PL, if any.
Digikrm spectrometer and an Optronis streak camera. The Fig. 2 shows the 300 K PL spectrum of the CSNWs and bare
measured TRPL resolution is  10 ps. All measurements were GaAs NWs. The sharp peaks in the 300 K spectra may be Ar laser
performed on the as-grown CSNW sample (i.e., on ensemble of plasma lines. For the CSNWs, we identied distinct bands at
CSNWs) only. TRPL for bare GaAs NWs was not reported because 1.437 eV and 1.559 eV corresponding to the GaAs core and to the
the signal was too weak to be detected by our TRPL setup. Al0.1Ga0.9As shell transitions, respectively. The position of the GaAs
band is near the bandgap of bulk zincblende GaAs, suggesting that
our sample is predomintantly zincblende (ZB). To verify this
3. Results structure, we use x-ray diffraction and temperature dependent
PL [32]. The low Al mole fraction x 0.1 may be attributed to the
Fig. 1 shows a representative top view SEM micrograph of the different kinetics between 2-D lm and NW growth, as observed
sample. The CSNWs are tilted with respect to the horizontal due to by Chen et al. [30]. In VLS, the growth depends on the rate at
the o1 1 1 4 preferential growth of GaAs NWs [28]. Close which the adatoms diffuse along the NW sidewalls [30]. Compared
inspection of the CSNWs show that they have a tapered tip, as to the bare GaAs NWs, there is more than an order of magnitude
previously observed in Au-assisted GaAs NWs [29]. From various increase in the GaAs PL intensity for the CSNWs. The enhanced
SEM images, the diameter of the CSNWs is estimated to be GaAs PL signal in CSNWs may be due to (i) suppressed surface
recombination because of the passivation provided by the Al0.1
Ga0.9As shell [11,12,1519]; and (ii) increased number of recom-
bining electron-hole pairs via the shell to core CT [21].
The shell-to-core CT in our sample is observed via PLE at 77 K.
Fig. 3a shows the 77 K PLE spectrum of GaAs where the excitation is
varied from below to above the Al0.1Ga0.9As bandgap of 1.65 eV. An
abrupt increase in the PLE signal is observed at 1.65 eV which we
attribute to the CT. At excitations ho1.65 eV (Region I), the carriers
that contribute to the GaAs PLE in Region I are those optically excited
from the GaAs region only. At excitations h41.65 eV (Region II),
carriers optically excited from the Al0.1Ga0.9As can diffuse and
transfer to GaAs and can contribute to the PLE signal [33]. Since
the PLE spectrum follows absorption spectrum for CW excitation, we
expect the PLE spectrum to mimic the E bulk density of states. The
PLE spectrum follows the h dependence in Region I but deviates
drastically in Region II. The deviation can be explained by the added
absorption introduced by the Al0.1Ga0.9As. The CT mechanism
Fig. 1. Top view SEM micrograph of the coreshell nanowires (CSNWs) showing
and band-to-band recombination are schematically shown in
high density formation. The CSNWs are predominantly tilted with respect to the Fig. 3b. However, it should be noted that aside from these two,
substrate surface due to the o 1 1 1 4 preferential growth of the nanowires. recombination via defects (not included in Fig. 3b) is also possible.
M.H. Balgos et al. / Journal of Luminescence 155 (2014) 2731 29

Fig. 3. (a) GaAs core PLE spectrum at 77 K. Abrupt increase of PL at excitations above the Al0.1Ga0.9As bandgap indicates carrier transfer. (b) Mechanism for the shell to core
carrier transfer.

Fig. 4. Time resolved photoluminescence at 77 K and 300 K of the (a) GaAs core and (b) AlGaAs shell of the core-shell nanowires. Black solid lines are ts to the coupled rate
equation (Eqs. (1a) and (1b)). The recombination time constants rec and carrier transfer time constants CT obtained from the ts are included in the gure.

The n-doped GaAs cap is too thin to have a signicant effect on the carriers to the band edges. The second term represents the
PLE signal. radiative (R) and nonradiative (NR) recombination with time
We further investigated the CT mechanism using TRPL. Fig. 4a constant 1/rec 1/R 1/NR. The last term accounts for of the
and b shows the TRPL spectra of GaAs and Al0.1Ga0.9As, respec- CT mechanism. The term in brackets is to incorporate the nite
tively, at 300 K and 77 K. The spectra were obtained by taking a number of carriers Nmax from Al0.1Ga0.9As that can transfer.
horizontal segment in the streak image with height equal to the Physically, only carriers near the Al0.1Ga0.9AsGaAs interface can
full-width at half maximum of the PL signal. All spectra were tted readily transfer; those away from the interface have to initially
using a coupled differential rate equation model to represent the diffuse before they can undergo CT. Simultaneously, in momentum
temporal carrier evolution for the Al0.1Ga0.9As shell and GaAs core: space, these carriers thermalize to reach the band edges before
  they recombine. Since thermalization is a faster process (in the
d N c t N ex;c Nc N s N max  N CT
 1a order of femtoseconds [35]) than diffusion, then recombination is
dt rc rec;c CT N max
favored at longer timescales. In our model, when NCT reaches Nmax,
  the CT process stops (last term in Eqs. (1a) and (1b) goes to zero)
d N s t N ex;s Ns N s N max  NCT
  1b leaving recombination as the only carrier exit path in the shell and
dt rs rec;s CT Nmax
the optical excitation as the only source of carriers in the core.
where the subscripts s and c denote shell and core, respectively. The black solid lines in Fig. 4a and b are the least square ts of
N is the number of carriers, Nex is the initial number of excited Eqs. (1a) and (1b), respectively. It is important to note that r is
carriers due to the optical excitation, r is the rise time constant, xed at 10 ps equivalent to the resolution of our TRPL setup. This
rec is the recombination time constant, CT is the transfer time value is reasonable since the thermalization of the carriers,
constant, Nmax is the maximum number of carriers that can characterized by r is in the order of femtoseconds [34] and is
transfer, and NCT is the number of carriers that transferred. Eqs. beyond the resolution of our TRPL setup.
(1a) and (1b) are modied generationrecombination rate equa- For the GaAs core (Fig. 4a), the TRPL rise has a single rate. Since
tions for NWs [34]. Generation manifests as the TRPL rise while carriers from Al0.1Ga0.9As transfer to GaAs, we expect a double rise
recombination manifests as the TRPL decay. The rst term in the characterized by two different time constants one from the
right-hand-side of Eqs. (1a) and (1b) indicates addition of carriers optical excitation and one from the CT. The single TRPL rise may be
via optical excitation and the subsequent thermalization of the explained by the relative contribution to the number of carriers of
30 M.H. Balgos et al. / Journal of Luminescence 155 (2014) 2731

Fig. 5. (a) Integrated PL intensity versus inverse temperature of the AlGaAs shell. Activation energy Ea 31.4 meV corresponds to the ionization of traps within the AlGaAs.
(b) Temperature dependent GaAsAlGaAs PL intensity ratio showing an increase at  90 K.

the two processes. More carriers are produced via optical excita- the GaAsAl0.1Ga0.9As PL intensity ratio (PLIR) as a function of
tion, hence, the TRPL rise from the CT is swamped and is not temperature. Owing to the same growth conditions, the defects
observed. The TRPL of GaAs at both temperatures is characterized incorporated in GaAs and AlGaAs have minimal difference. The PL
by a single TRPL decay rate. We attribute the decay to carrier signal for GaAs and Al0.1Ga0.9As therefore should quench at almost
recombination within the GaAs, The faster decay of 600 ps at the same rate with temperature resulting to a constant PLIR if they
300 K is due to signicant NR recombination at 300 K. Since rec,c are to be treated independently. Constant PLIR is seen at 1070 K
which is a combined effect of R and NR recombination [36], that is, as indicated by the horizontal dash line in Fig. 5b. At 90150 K,
1/rec 1/R 1/NR, we observe a longer time constant of 970 ps at however, there is an apparent increase in the PLIR as indicated by
77 K where NR is minimized and 1/NR approaches zero. The the slanted dash line. At these temperatures, the GaAs PL intensity
values obtained at 300 K and 77 K are within the range of reported increases relatively to the Al0.1Ga0.9As. More carriers from the
carrier lifetime values for nearly intrinsic exciton lifetimes of GaAs Al0.1Ga0.9As transferred to GaAs at T 490 K. We note that
NWs (0.051.1 ns) reported by Perera et al. [37]. the intersection of the horizontal and slanted line at  90K
For the Al0.1Ga0.9As shell (Fig. 4b), a single exponential TRPL coincides with the activation temperature of the defects within
rise, due to the optical excitation, is observed at 77 K and 300 K. the Al0.1Ga0.9As.
The TRPL decay, however, is characterized by a single exponential
decay rate at 77 K and a double exponential (fast and slow) decay
rate at 300 K. We propose that the appearance of the double decay 4. Conclusion
is a consequence of the CT mechanism. We attribute the fast decay
to the combined effect of recombination and carrier transfer while The shell-to-core CT mechanism of an ensemble of dominantly
the slow decay is ascribed to the recombination. Numerical tting zincblende GaAs/AlGaAs CSNWs is investigated through CW-PL,
using Eq. (1a) revealed rec,s of 400 ps at 300 K and 580 ps at 77 K. TRPL and PLE spectroscopy for varying temperatures. PLE provides
A longer rec,s at 77 K is again attributed to the minimized NR strong evidence of the CT mechanism while CW-PL and TRPL
recombination at low temperatures. The transfer time constant CT demonstrate that the CT process is dependent of temperature
is 50 ps and 55 ps for at 300 K and 77 K, respectively. This is mainly because a high number of carriers participate in the CT
similar to the CT time constants of GaAs/AlGaAs separate conne- process at T 490 K. Coupled rate equations are used to t the TRPL
ment heterostructures reported by Morin et al. (19.2 ps at 300 K at 300 K and at 77 K, thereby facilitating the evaluation of the
and 22 ps at 80 K) [38]. Our values of CT show that the CT time is recombination and carrier transfer time constants. The 300 K
almost independent of temperature and cannot explain the double (77 K) average recombination time constants for GaAs core are
decay observed in the 300 K TRPL of Al0.1Ga0.9As. rec,c 600 ps (970 ps) while that of the AlGaAs shell are rec,
As such, the the maximum number of transferred carriers Nmax s 400 ps (580 ps). The CT time constant is determined to be

relative to the initial number of carriers Nex,s (t0) in the shell was CT 50 ps (55 ps) for 300 K (77 K).
investigated. We obtain a value of Nmax/Nex,s(t0) of 0.24 at 77 K
and 0.69 at 300 K. These values show that a higher fraction of
Acknowledgments
excited carriers transfer at 300K than at 77 K. We believe that
defects in the Al0.1Ga0.9As affect the number of carriers that can
This work is supported in part by Grants from PCIEERD-DOST,
transfer. Using an Arrhenius plot of the integrated PL intensity vs
DOST-GIA, and University of the Philippines OVCRD. The authors
inverse temperature (Fig. 5a), we calculate an activation energy
would also like to thank Intel for the donation of the streak
Ea 31 meV for the defects within the AlGaAs at  90 K which can
camera.
be attributed to an electron trap E9 (30 meV) or a hole trap H4
(32 meV) [39]. At 77 K (o90 K), there is still efcient trapping of
carrier by the defects, thus there are not enough transferred References
carriers to produce a double decay. At 300 K, the defects in the
[1] B.R. Huang, Y.K. Yang, T.C. Lin, W.L. Yang, Sol. Energy Mater. Sol. Cells 98 (2012)
shell are already ionized and are not as efcient at trapping
357.
carriers. Thus, more carriers can diffuse until they nally transfer [2] Y. Liang, H. Liang, X. Xiao, S. Hark, J. Mater. Chem. 22 (2012) 1199.
to the GaAs. [3] S.L. Wu, J.H. Deng, T. Zhang, R.T. Zheng, G.A. Cheng, Diam. Relat. Mater.
The temperature dependence of the number of carriers avail- 26 (2012) 83.
[4] T. Mrtensson, C.P.T. Svensson, B.A. Wacaser, M.W. Larsson, W. Seifert,
able for transfer to GaAs is investigated using the relative PL K. Deppert, A. Gustafsson, L.R. Wallenberg, L. Samuelson, Nano Lett. 4 (2004)
intensities of GaAs and Al0.1Ga0.9As at 10K300 K. Fig. 5b shows 1987.
M.H. Balgos et al. / Journal of Luminescence 155 (2014) 2731 31

[5] K. Tomioka, T. Tanaka, S. Hara, K. Hiruma, T. Fukui, IEEE J. Sel. Top. Quantum [24] B. Deveaud, F. Clerot, A. Regreny, K. Fujiwara, K. Mitsunaga, J. Ohta., Appl. Phys.
Electron. 17 (2011) 1112. Lett. 55 (25) (1989) 2646.
[6] G.E. Cirlin, V.G. Dubrovskii, Y.B. Samsonenko, A.D. Bouravleuv, K. Durose, [25] R.S. Wagner, W.C. Ellis, Appl. Phys. Lett. 4 (5) (1964) 89.
Y.Y. Proskuryakov, B. Mendes, L. Bowen, M.A. Kaliteevski, M.A. Kaliteevski, [26] H.L. Hartnagel, R. Riemenschneider, Properties of Aluminium Gallium
D. Zeze, Phys. Rev. B 82 (2010) 035302. Arsenide, in: S. Adachi (Ed.), INSPEC, the Institution of Electrical Engineers,
[7] L.C. Chuang, M. Moewe, S. Crankshaw, C. Chang-Hasnain, Appl. Phys. Lett. 92 London, United Kingdom, 1993, p. 239.
(2008) 013121. [27] J.P. Callan, A.M.T. Kim, L. Huang, E. Mazur, Chem. Phys. 251 (2000) 167.
[8] J. Bao, M.A. Zimmler, F. Capasso, X. Wang, Z.F. Ren, Nano Lett. 6 (2006) 1719. [28] Seth A. Fortuna, Li Xiuling, Semicond. Sci. Technol. 25 (2) (2010) 024005.
[9] R. Konenkamp, R.C. Word, C. Schlegel, Appl. Phys. Lett. 85 (2004) 6004. [29] M.C. Plante, R.R. LaPierre, Au-assisted growth of GaAs nanowires by gas source
[10] S. Gradeak, F. Qian, Y. Li, H.G. Park, C.M. Lieber, Appl. Phys. Lett. 87 (2005) molecular beam epitaxy: tapering, sidewall faceting and crystal structure,
173111. J. Cryst. Growth 310 (2) (2008) 356.
[11] O. Demichel, M. Heiss, J. Bleuse, H. Mariette, A. Fontcuberta i Morral, Appl. [30] Chen Chen, Shyemaa Shehata, Ccile Fradin, Ray LaPierre, Christophe Couteau,
Phys. Lett. 97 (2010) 201907. Weihs Gregor, Nano Lett. 7 (no. 9) (2007) 2584.
[12] C.Y. Chen, C.A. Lin, M.J. Chen, G.R. Lin, J.H. He, Nanotechnology 20 (2009) [31] Yu B. Bolkhovityanov, Pchelyakov Oleg Petrovich, Phys. Uspekhi 51 (5) (2008)
437.
185605.
[32] Michelle F. Bailon-Somintac, Jasher J. Ibaez, Rafael B. Jaculbia, Regine
[13] X. Jiang, Q. Xiong, S. Nam, F. Qian, Y. Li, C.M. Lieber, Nano Lett. 7 (2007) 3214.
A. Loberternos, Michael J. Defensor, Arnel A. Salvador, Armando S. Somintac,
[14] P. Parkinson, H.J. Joyce, Q. Gao, H.H. Tan, X. Zhang, J. Zou, C. Jagadish, L.M. Herz,
Low temperature photoluminescence and Raman phonon modes of Au-
M.B. Johnston, Nano Lett. 9 (2009) 3349.
catalyzed MBE-grown GaAsAlGaAs coreshell nanowires grown on a pre-
[15] J. Noborisaka, J. Motohisa, S. Hara, T. Fukui, Appl. Phys. Lett. 87 (2005) 093109.
patterned Si (111) substrate, J. Cryst. Growth 314 (1) (2011) 268.
[16] N. Skld, L.S. Karlsson, M.W. Larsson, M.E. Pistol, W. Seifert, J. Trgrdh,
[33] B. Ketterer, M. Heiss, M.J. Livrozet, A. Rudolph, E. Reiger, A. Fontcuberta i
L. Samuelson, Nano Lett. 5 (2005) 1943.
Morral, Phys. Rev. B 83 (2011) 125307.
[17] P. Prete, F. Marzo, P. Paiano, N. Lovergine, G. Salviati, L. Lazzarini, T. Sekiguchi,
[34] L. Smith, H. Jackson, J. Yarrison-Rice, C. Jagadish, Semicond. Sci. Technol. 25
J. Cryst. Growth 310 (2008) 5114. (2010) 024010.
[18] O.D.D. Couto Jr., D. Sercombe, J. Puebla, L. Otubo, I.J. Luxmoore, M. Sich, [35] J.P. Callan, A.M-T. Kim, C.A.D. Roeser, E. Mazur, Semiconductors and Semi-
T.J. Elliott, E.A. Chekhovich, L.R. Wilson, M.S. Skolnick, H.Y. Liu, A. metals, in: K.T. Tsen (Ed.), Elsevier B.V., Massachusetts, USA, 2001,
I. Tartakovskii, Nano Lett. 12 (2012) 5269. pp. 151203.
[19] C. Chen, S. Shehata, C. Fradin, R. LaPierre, C. Couteau, G. Weihs, Nano Lett. 7 [36] P.Y. Yu, M. Cardona, Fundamentals of Semiconductors: Physics and Materials
(2007) 2584. Properties, fourth ed., Springer-Verlag, Berlin Heidelberg, New York, 2010.
[20] C.H. Chuang, T. Doane, S. Lo, G. Scholes, C. Burda., Measuring electron and hole [37] S. Perera, M.A. Fickenscher, H.E. Jackson, L.M. Smith, J.M. Yarrison-Rice,
transfer in core/shell nanoheterostructures, ACS Nano 5 (7) (2011) 6016. H.J. Joyce, Q. Gao, H.H. Tan, C. Jagadish, X. Zhang, J. Zou, Appl. Phys. Lett.
[21] J. Yoo, B. Chon, W. Tang, T. Joo, L.S. Dang, G.C. Yi, Appl. Phys. Lett. 100 (2012) 93 (2008) 053110.
223103. [38] S. Morin, B. Deveaud, F. Clerot, K. Fujiwara, K. Mitsunaga, IEEE J. Quantum
[22] B. Deveaud, J. Shah, T.C. Damen, W.T. Tsang., Appl. Phys. Lett. 52 (22) (1988) Electron. 27 (1991) 1669.
1886. [39] Sheinkman, T.V. Torchinskayaand M.K., Deep Centers In AlGaAs Heteroepitaxial
[23] J.A. Brum, G. Bastard, Phys. Rev. B 33 (2) (1986) 1420. Diodes.

Вам также может понравиться