Você está na página 1de 8

Time-resolved photoluminescence of silicon

microstructures fabricated by femtosecond laser


in air
Zhandong Chen,1 Qiang Wu,1,* Ming Yang,1 Jianghong Yao,1 Romano A. Rupp,1,2
Yaan Cao,1 and Jingjun Xu1
1
The MOE Key Laboratory of Weak Light Nonlinear Photonics, TEDA Applied Physics School and School of
Physics, Nankai University, Tianjin 300457, China
2
Vienna University, Faculty of Physics, Strudlhofstiege 4, A-1090 Wien, Austria
*wuqiang@nankai.edu.cn

Abstract: Green photoluminescence (PL) from silicon microstructures


fabricated by femtosecond laser in air was studied at different temperature
by time-resolved spectroscopy. The PL decay profiles are well fitted by a
stretched exponential function: I (t ) = I (0) ∗ exp[−(t / τ ) β ]. The dependence
of the decay time constant τ and of the stretching index β on PL photon
energy and on the temperature is investigated. A model of transport and
recombination of the carriers is introduced as a possible explanation of the
stretched exponential decay. The nonradiative recombination rate of the
localized carriers, which is dependent on the carrier density and influenced
by the trapping site density and the temperature, is deduced to be
responsible for this kind of decay.
©2013 Optical Society of America
OCIS codes: (300.6500) Spectroscopy, time-resolved; (250.5230) Photoluminescence;
(320.7130) Ultrafast processes in condensed matter, including semiconductors.

References and links


1. L. T. Canham, “Silicon quantum wire array fabrication by electrochemical and chemical dissolution of wafer,”
Appl. Phys. Lett. 57(10), 1046–1048 (1990).
2. V. Lehmann and U. Gosele, “Porous silicon formation: A quantum wire effect,” Appl. Phys. Lett. 58(8), 856–
858 (1991).
3. T. Shimizu-lwayama, S. Nakao, and K. Saitoh, “Visible photoluminescence in Si+-implanted thermal oxide films
on crystalline Si,” Appl. Phys. Lett. 65(14), 1814–1816 (1994).
4. R. J. Walters, G. I. Bourianoff, and H. A. Atwater, “Field-effect electroluminescence in silicon nanocrystals,”
Nat. Mater. 4(2), 143–146 (2005).
5. K. Žídek, F. Trojánek, P. Malý, L. Ondič, I. Pelant, K. Dohnalová, L. Šiller, R. Little, and B. R. Horrocks,
“Femtosecond luminescence spectroscopy of core states in silicon nanocrystals,” Opt. Express 18(24), 25241–
25249 (2010).
6. X. Chen, D. Uttamchandani, C. Trager-Cowan, and K. P. O’Donnell, “Luminescence from porous silicon,”
Semicond. Sci. Technol. 8(1), 92–96 (1993).
7. M. Zhu, Y. Han, R. B. Wehrspohn, C. Godet, R. Etemadi, and D. Ballutaud, “The origin of visible
photoluminescence from silicon oxide thin films prepared by dual-plasma chemical vapor deposition,” J. Appl.
Phys. 83(10), 5386–5393 (1998).
8. T. Schmidt, A. I. Chizhik, A. M. Chizhik, K. Potrick, A. J. Meixner, and F. Huisken, “Radiative exciton
recombination and defect luminescence observed in single silicon nanocrystals,” Phys. Rev. B 86(12), 125302
(2012).
9. K. Kůsová, O. Cibulka, K. Dohnalová, I. Pelant, J. Valenta, A. Fucíková, K. Zídek, J. Lang, J. Englich, P.
Matejka, P. Stepánek, and S. Bakardjieva, “Brightly luminescent organically capped silicon nanocrystals
fabricated at room temperature and atmospheric pressure,” ACS Nano 4(8), 4495–4504 (2010).
10. D. S. English, L. E. Pell, Z. Yu, P. F. Barbara, and B. A. Korgel, “Size tunable visible luminescence from
individual organic monolayer stabilized silicon nanocrytal quantum dots,” Nano Lett. 2(7), 681–685 (2002).
11. L. Tsybeskov, J. V. Vandyshev, and P. M. Fauchet, “Blue emission in porous silicon: Oxygen-related
photoluminescence,” Phys. Rev. B Condens. Matter 49(11), 7821–7824 (1994).
12. H. Tamura, M. Ruckschloss, T. Wirschem, and S. Veprek, “Origin of the green/blue luminescence from
nanocrystalline silicon,” Appl. Phys. Lett. 65(12), 1537–1539 (1994).
13. G. Ledoux, J. Gong, and F. Huisken, “Effect of passivation and aging on the photoluminescence of silicon
nanocrystals,” Appl. Phys. Lett. 79(24), 4028–4030 (2001).

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21329
14. V. Vinciguerra, G. Franzo, F. Priolo, F. Iacona, and C. Spinella, “Quantum confinement and recombination
dynamics in silicon nanocrystals embedded in Si/SiO2 superlattices,” J. Appl. Phys. 87(11), 8165–8173 (2000).
15. C. Wu, C. H. Crouch, L. Zhao, and E. Mazur, “Visible luminescence from silicon surfaces microstructured in
air,” Appl. Phys. Lett. 81(11), 1999–2001 (2002).
16. C. Wu, C. H. Crouch, L. Zhao, J. E. Carey, R. Younkin, J. A. Levinson, E. Mazur, R. M. Farrell, P. Gothoskar,
and A. Karger, “Near-unity below-band-gap absorption by microstructured silicon,” Appl. Phys. Lett. 78(13),
1850–1852 (2001).
17. M. Y. Shen, C. H. Crouch, J. E. Carey, R. Younkin, E. Mazur, M. Sheehy, and C. M. Friend, “Formation of
regular arrays of silicon microspikes by femtosecond laser irradiation through a mask,” Appl. Phys. Lett. 82(11),
1715–1717 (2003).
18. J. E. Carey, C. H. Crouch, M. Y. Shen, and E. Mazur, “Visible and near-infrared responsivity of femtosecond-
laser microstructured silicon photodiodes,” Opt. Lett. 30(14), 1773–1775 (2005).
19. Z. H. Huang, J. E. Carey, M. G. Liu, X. Y. Guo, E. Mazur, and J. C. Campbell, “Microstructured silicon
photodetector,” Appl. Phys. Lett. 89(3), 033506 (2006).
20. Q. Wu, S. Guo, Y. Ma, F. Gao, C. Yang, M. Yang, X. Yu, X. Zhang, R. A. Rupp, and J. Xu, “Optical refocusing
three-dimensional wide-field fluorescence lifetime imaging microscopy,” Opt. Express 20(2), 960–965 (2012).
21. A. Menéndez-Manjón, S. Barcikowski, G. A. Shafeev, V. I. Mazhukin, and B. N. Chichkov, “Influence of beam
intensity profile on the aerodynamic particle size distributions generated by femtosecond laser ablation,” Laser
Part. Beams 28(01), 45–52 (2010).
22. Y. L. Wang, C. Chen, X. C. Ding, L. Z. Chu, Z. C. Deng, W. H. Liang, J. Z. Chen, and G. S. Fu, “Nucleation
and growth of nanoparticles during pulsed laser deposition in an ambient gas,” Laser Part. Beams 29(01), 105–
111 (2011).
23. S. Manickam, K. Venkatakrishnan, B. Tan, and V. Venkataramanan, “Study of silicon nanofibrous structure
formed by femtosecond laser irradiation in air,” Opt. Express 17(16), 13869–13874 (2009).
24. Z. Chen, Q. Wu, M. Yang, B. Tang, J. Yao, R. A. Rupp, Y. Cao, and J. Xu, “Generation and evolution of plasma
during femtosecond laser ablation of silicon in different ambient gases,” Laser Part. Beams (to be published).
25. M. V. Wolkin, J. Jorne, P. M. Fauchet, G. Allan, and C. Delerue, “Electronic states and luminescence in porous
silicon quantum dots: The role of oxygen,” Phys. Rev. Lett. 82(1), 197–200 (1999).
26. J. Martin, F. Cichos, F. Huisken, and C. von Borczyskowski, “Electron-phonon coupling and localization of
excitons in single silicon nanocrystals,” Nano Lett. 8(2), 656–660 (2008).
27. R. W. Collins, M. A. Paesler, and W. Paul, “The temperature dependence of photoluminescence in a-Si: H
alloys,” Solid State Commun. 34(10), 833–836 (1980).
28. R. Kohlrausch, “Nachtrag ueber die elastische Nachwirkung beim Cocon-und Glasfaden, und die
hygroskopische Eigenschaft des ersteren,” Ann. Phys. (Leipzig) 12, 393–399 (1847).
29. B. Sturman, E. Podivilov, and M. Gorkunov, “Origin of stretched exponential relaxation for hopping-transport
models,” Phys. Rev. Lett. 91(17), 176602 (2003).
30. T. Bartel, M. Dworzak, M. Strassburg, A. Hoffmann, A. Strittmatter, and D. Bimberg, “Recombination
dynamics of localized excitons in InGaN quantum dots,” Appl. Phys. Lett. 85(11), 1946–1948 (2004).
31. M. Dovrat, Y. Goshen, J. Jedrzejewski, I. Balberg, and A. Sa’ar, “Radiative versus nonradiative decay processes
in silicon nanocrystals probed by time-resolved photoluminescence spectroscopy,” Phys. Rev. B 69(15), 155311
(2004).
32. S. E. Paje and J. Llopis, “Photoluminescence decay and time-resolved spectroscopy of cubic yttria-stabilized
zirconia,” Appl. Phys., A Mater. Sci. Process. 59(6), 569–574 (1994).
33. F. Sangghaleh, B. Bruhn, T. Schmidt, and J. Linnros, “Exciton lifetime measurements on single silicon quantum
dots,” Nanotechnology 24(22), 225204 (2013).

1. Introduction
In past decades, much effort has been devoted to achieve light-emitting devices based on
silicon, in order to conveniently integrate electronic and optical functions onto the same chip.
However, the nature of the indirect band gap of silicon makes it an inefficient light source.
Several approaches were undertaken to overcome this difficulty, e.g., using porous silicon
(PS) [1,2] or silicon nanocrystals (Si NCs) [3–5]. Different PL bands have been observed on
various samples of PS and Si NCs [6–10]. A blue-green band has been attributed to surface
localized states related to oxygen defects [11,12]. A red band, whose peak wavelength has a
blue shift as the size of silicon nanocrystals decreases, originates from a quantum
confinement effect [13,14]. Recently, photoluminescence (PL) was observed at room
temperature on microstructured silicon fabricated in air by a femtosecond laser [15]. It is
promising to be used for the light-emitting devices. Moreover, femtosecond laser
microstructured silicon has a very high absorptance from the near ultraviolet to the near
inferred, which makes it possible to be used to improve the efficiency of silicon solar cells
[16]. This material exhibits advantages in the field of field-emission and photo-detection as
well [17,18]. It is notable that the photo-detector based on femtosecond laser microstructured
silicon has exhibited excellent performance [19]. Up to now, however, the dynamics of photo-

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21330
generated carriers in microstructured silicon, which is very important information for
improving the performance of the devices based on this kind of material, are still unclear, due
to the complexity of the material.
In this letter, PL of the microstructured silicon was studied at different temperature by
time-resolved spectroscopy that is a powerful tool to study the mechanism of PL and the
dynamics of the carriers in a complex system [20], aiming to solve the problems mentioned
above. The temperature dependence of PL intensity implies a wide band tail in the
microstructured silicon. The PL decay profile is nonexponential, which is well fitted with a
stretched exponential function. The dependences of the fitting parameters (the decay time
constant and the stretching index) on temperature and on PL photon energy were determined.
To explain the stretched exponential decay of PL and illustrate the dynamics of the carriers, a
model of transport and recombination of the carriers is established. The effect of the trapping
sites is discussed as well. This model is well consistent with the experimental results.
2. Experiments
The sample was fabricated by irradiating an n-type silicon (111) wafer in air with a train of
120 fs laser pulses at a repetition rate of 1 kHz and at a central wavelength of 800 nm. The
silicon wafers, which had a resistivity of more than 1000 Ωcm , were cleaned with
hydrofluoric acid to remove any native oxide and then rinsed with distilled water.
Microstructured silicon patches (5 mm × 5 mm) were produced by translating the silicon
wafer during irradiation under a laser fluence of 10 kJ/m2 in air. Subsequently, the sample
was annealed at 1300 K for 1 hour in vacuum.
PL investigations were performed with the experimental setup shown in Fig. 1. The 800
nm femtosecond laser was frequency doubled to 400 nm using a BBO crystal to excite the
sample that was placed in a hot and cold stage (HCS402, INSTEC). This stage can keep the
sample temperature from 90 K to 300 K. PL signals were collected by a light collection
system, which contained a longpass filter to block the scattered pump light. The collected
signal was focused onto the slit of a spectrograph (SpectraPro-300i, Acton Research
Corporation) coupled to an ICCD camera (PicoStar HR 12, LaVision). The ICCD and the
femtosecond laser system were synchronized with a precision better than 10 ps. A
programmable delay generator was used to control the delay between the laser pulse and the
gate signal of the ICCD. The minimum width of the gate signal is 30 ps.

Fig. 1. The schematic of the experimental setup. SHG: second harmonic generation; Filter1:
blue bandpass filter; M1: reflection mirror; L1, L2, L3: lens; Filter2: longpass filter.

3. Results and discussion


The inset [Fig. 2(b)] shows the SEM (scanning electron microscopy) image of the surface of
the annealed sample. After femtosecond laser ablation in air, micro-cones with a height of

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21331
some tens of microns are seen on the silicon surface. The cones are covered by smaller
microstructures that contain silicon nanocrystals with an oxide layer [15]. This kind of
generation of nanoparticles and nanostructures is very common during femtosecond laser
ablation [21–23]. The removed material in the ablated plume can redeposit on the surface of
the sample due to confinement of the ambient gas [24], thus forming the nanostructures.
The PL spectra shown in Fig. 2 are obtained by temporally integrating the time-resolved
spectra measured at 90 K and 300 K respectively. The peak wavelength, which is
approximately at 530 nm for both annealed sample and unannealed sample, does not vary
with sample temperature. This green band is due to oxygen-related defects that locate at the
Si/SiO2 interface [15] and can greatly affect the properties of the silicon nanocrystals [25,26].
This can be confirmed by the disappearance of the green band after removing the oxide layer
on the sample’s surface. As shown in Fig. 2, the PL intensity of the annealed sample is higher
than that of the unannealed sample due to the decrease of the trapping sites (nonradiative
recombination centers) after annealing. Hence, we paid more attention to the annealed sample
in this study.

Fig. 2. Temporal integral of time-resolved PL spectra of microstructured silicon obtained at 90


K and 300 K, respectively. The inset (a) shows the temperature dependence of the PL intensity
of the annealed sample at peak wavelength. The solid line represents a theoretical fit. The inset
(b) shows the SEM image of the annealed sample.

As seen in Fig. 2, the PL intensity is much higher at 90 K than that at 300 K. To determine
the temperature effect, the peak intensity of the annealed sample is plotted as a function of
temperature, as shown in Fig. 2(a). From 90 K to 300K, the PL intensity decreases with
temperature. This is due to the competition between the radiative and nonradiative
recombination. At low temperature, the photogenerated carriers are more easily localized in
the surface states that lie in the band tails and are related with the oxygen-related defects at
Si/SiO2 interface, causing a stronger optical radiation. As the temperature increases, lots of
carriers can be reemitted from the surface states to the mobility edge and diffuse to the
trapping sites, and then recombine nonradiatively, leading to a decrease of the optical
emission. The dependence of PL intensity on temperature can be described by [6]:
I = I 0 / [1 + B ∗ exp(T / T0 )]. (1)
Here B and I0 are constants, i.e., independent of temperature, and T0 denotes the degree of
the system disorder, related to the tail width of the density of localized states [27]. From the
fit we obtain T0~163 K (i.e. kBT0~14 meV). This points to a wide band tail in the
microstructured silicon.

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21332
Figures 3(a) and 3(c) show the time-resolved PL spectra of the annealed sample at 90 K
and 300 K. The PL decay profiles at different wavelengths are extracted from them, as shown
in Figs. 3(b) and 3(d). It is evident that the decay processes are temperature dependent and
that the PL decay profiles are nonexponential. The decay profiles are well fitted by a stretched
exponential function [28]:
I (t ) = I (0) ∗ exp[−(t / τ ) β ]. (2)
where τ is a time constant, and 0 ≤ β ≤ 1 is the stretching index. The stretched exponential
function is well known to describe the PL decay and transport properties of disordered
systems [29–31]. The stretching index is a measure for the degree of disorder in the material.

Fig. 3. Time-resolved PL spectra of the annealed sample measured at (a) 300 K and (c) 90 K.
The decay profiles at different wavelengths are obtained at (b) 300 K and (d) 90 K. The solid
lines are the fits with the stretched exponential function.

As seen in Figs. 3(b) and 3(d), the PL decay profiles at different wavelengths differ for
300 K and 90 K. The fitting results are listed in Table 1.
Table 1. Values of τ and β Obtained by Fitting PL Decay Data with Eq. (2) at Different
Wavelengths at 300 K and 90 K.

WL (nm) 530 590 620 650

T (K) 300 90 300 90 300 90 300 90

τ (ns) 1.43 2.41 1.23 2.03 0.87 1.68 0.66 1.16

β 0.58 0.62 0.60 0.63 0.58 0.64 0.59 0.62

At a certain temperature, the time constant τ decreases with wavelength while β nearly
keeps constant. This suggests that the surface states with different energy levels have different
lifetimes. The invariable β proves that the mechanisms causing the stretched exponential
decay are similar at different wavelengths. We will introduce a model to explain the
mechanisms in more detail in next paragraph. It is notable that the values of τ and β are larger
at 90 K than that at 300 K. To determine the dependence of τ and β on temperature, the values
of τ and β for the wavelength of 530 nm are plotted as a function of temperature, as shown in
Fig. 4. The value of β varies only a little from 90 K to 240 K, while visibly decreases for
temperatures above 240 K. The value of τ decreases with temperature, which can be

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21333
explained by an enhancement of the nonradiative recombination at higher temperature. The
temperature dependence of the time constant τ can be described by the classical expression
[32]:
τ = τ 0 / [1 + s ∗ exp(− E / k BT )]. (3)
where τ0 and s are constants, and E denotes the activation energy. Applying Eq. (3) yields an
activation energy of E~98.6 meV.

Fig. 4. Dependences of the time constant τ and the stretching index β on temperature for
the wavelength of 530 nm. The data of the time constant are fitted by using Eq. (3) and the
solid curve through the data of the stretching index serves only to guide the eye.

The decay of PL of the microstructured silicon is well described by the stretched


exponential function. Many efforts have been made to find out the physics behind the
exponential decay [29, 33]. However, the physical mechanism is unclear yet. For this sake,
we introduce a model of transport and recombination of the carriers as a possible explanation
for the stretched exponential decay. When a pump laser pulse impacts upon the sample
surface, large amounts of carriers are excited in the cores of the silicon nanocrystals, and then
quickly transmit to the Si/SiO2 interface and are localized in the surface states. The localized
carriers can recombine radiatively, or be thermally reemitted from the localized states. The
reemitted carriers may diffuse to the neighboring trapping sites and recombine nonradiatively,
or be re-localized by the surface states. During the PL process, the radiative recombination
and the total nonradiative recombination compete with each other. The dynamics of
reemission, diffusion, re-localizing, trapping, and nonradiative recombination of the localized
carriers, which is dependent on the carrier density and influenced by the trapping site density
and the temperature, determines the nonexponential decay of PL.
By using a diffusion model, we estimate that the carriers transmit from the silicon core
(several nanometers) to the Si/SiO2 interface on the time scale of tens to hundreds of
picosecond, which is much faster than the decay of PL. Hence, we approximately consider the
decay of PL with an initial density of the localized carriers. The density of the localized
carriers can be described as:
dN dt = − Asr ∗ N − Asnr ∗ N . (4)
where Asr is the radiative recombination rate, which is a constant. Asnr is the nonradiative
recombination rate, which is determined by the dynamics of the localized carriers and should
be a function of the carrier density. We assume that Asnr can be described as Asnr = Asnr 0 ∗ N b .

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21334
Here, the coefficient Asnr0 and the exponent b are constant. Hence, the optical emission
intensity can be described by:
I (t ) = ω ∗ Asr ∗ N (t ). (5)

where ω denotes the photon energy. We estimate the initial density (N0) of the excited
carriers at ~4 × 1018 cm−3 by considering the incident laser fluence of 0.4 mJ/cm2 (used in our
experiment) and a measured absorptivity of 0.9 for the laser at 400 nm. The decay profiles of
PL at 530 nm for the annealed sample are fitted by Eq. (5) combined with Eq. (4), as shown
in Fig. 5. The experimental results are perfectly consistent with this model. This hints that the
stretched exponential decay of PL is possibly caused by the nonradiative recombination rate
that is dependent on the carrier density.

Fig. 5. The decay profiles of the PL (annealed sample) at 530 nm under 300 K and 90 K
respectively. The points represent the experimental data and the solid lines denote the fits with
Eq. (5). The fitting parameters are listed in the inset table.

As seen in the inset table of Fig. 5, the coefficient Asnr0 is larger at higher temperature. It
means that the nonradiative recombination is more intense at higher temperature due to a
stronger reemission of localized carriers and an enhancement of nonradiative recombination,
leading to a smaller β. Besides dependence on sample’s temperature, the decay profile is
influenced by the density of trapping sites as well. We obtain Asnr0~6.9 × 10−10 and β~0.42 for
an unannealed sample at 300 K, while they are 2.4 × 10−10 and 0.58 for the annealed sample
respectively. This is consistent with the fact that the trapping sites remarkably decrease after
high-temperature annealing. As a result, it is more difficult for the reemitted carriers to
diffuse to the trapping site and nonradiatively recombine, which suppresses the nonradiative
recombination.
4. Summary
In summary, PL of silicon microstructures fabricated by femtosecond laser in air has been
studied at different temperature by time-resolved spectroscopy. The stretched exponential
decay of PL implies a complex dynamics of photo-generated carriers in microstructured
silicon. The intensity and the decay time of PL decrease with temperature, suggesting a
competition between radiative and nonradiative recombination. This competition is governed
by the carrier dynamics. A model of transport and recombination of the carriers is introduced
to illustrate the carrier dynamics and explain the stretched exponential decay. The
nonradiative recombination rate, which is determined by a complex process of reemission,
diffusion, re-localizing, trapping, and nonradiative recombination of the localized carriers, is

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21335
dependent on the carrier density and influenced by the trapping site density and the
temperature. This is deduced to be responsible for the stretched exponential decay. The effect
of the trapping sites should be similar in other photoelectric devices based on the
microstructured silicon. Our results are helpful to analyze the dynamics of the carriers in
these devices as well.
Acknowledgments
This work was supported by the National Basic Research Program of China (2012CB934201
and 2013CB328702), the 111 Project (B07013), and the National Natural Science Foundation
of China (11074129).

#193816 - $15.00 USD Received 15 Jul 2013; revised 28 Aug 2013; accepted 28 Aug 2013; published 4 Sep 2013
(C) 2013 OSA 9 September 2013 | Vol. 21, No. 18 | DOI:10.1364/OE.21.021329 | OPTICS EXPRESS 21336

Você também pode gostar