Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Near-infrared anti-Stokes photoluminescence of PbS QDs embedded in glasses

Open Access Open Access

Abstract

Near-infrared photoluminescence properties of PbS QDs embedded in glasses were investigated upon below-bandgap excitation. PbS QDs were precipitated in the glasses upon thermal treatment. Near-infrared anti-Stokes photoluminescence (ASPL) from PbS QDs was observed. Dependence of the ASPL on size and excitation power indicated that ASPL was phonon-assisted one-photon process. These near-infrared anti-Stokes photoluminescence of PbS QDs in glasses have potential applications for light conversion and laser cooling.

© 2017 Optical Society of America

1. Introduction

Anti-Stokes photoluminescence is luminescence where the emitted photons are higher in energy than the excitation photons. ASPL, which violates of Stokes’ law in nature, was first observed in dye molecules [1], and then has been reported in various systems like bulk semiconductor and nanocrystals [2–4], polymers [5,6], and heterostructures [7,8] Due to the particular properties, ASPL has great potentials for applications in the multi-color displays [9], dynamical imaging microscopy [10], unconventional lasers [11], bio-imaging agents [12] and solid-state laser cooling devices [13].

Two-photon absorption process [3,4], Auger recombination [7,8], and one-photon absorption assisted by phonons [14–19] have been considered as the possible mechanisms leading to ASPL. In the two-photon absorption process, two photons are absorbed through a real or virtual intermediate state leading to higher energy recombination. Two photons absorption through a real intermediate state is common in ASPL for rare-earth ions while through a virtual intermediate state has been used to explain the ASPL in semiconductor nanocrystals [3,4]. The second process is an Auger-like recombination, where an electron-hole pair recombines nonradiatively and excites another electron or hole into higher energy states, which is common in heterostructures [7,8]. The third process involves one-photon phonon-assisted carrier excitation. The additional energy comes from the absorption of phonons, which has been reported to explain ASPL of QDs, like CdSe [14,15], InP [15], CdTe [14,16], PbS [17] and PbSe [18,19]. In general, intensity of ASPL (IASPL) based on the two-photon absorption shows a quadratic dependence on the excitation power (Iex) [20]. For the ASPL induced by the Auger recombination, a cubic dependence on the excitation power should be observed [20], and for the phonon absorption assisted ASPL, it is linearly dependent on the excitation power [15,16,20].

Synthesis of quantum dots in glasses is particularly attractive since it is inexpensive and chemically durable with high mechanical strength [21]. Furthermore, the versatility of glassy hosts allows fabrication of devices in the bulk, planar or fiber forms [21]. However, there has been not much researches on the ASPL of QDs doped glasses.

In this work, a series of PbS QDs with different diameters were precipitated inside the silicate glasses. Upon excitation at the long wavelength side of the absorption bands, efficient anti-Stokes photoluminescence was observed. The anti-Stokes photoluminescence spectrum was significantly red shifted relative to the normal photoluminescence spectrum. Moreover, with the increase in the excitation intensity, center wavelength of the anti-Stokes photoluminescence band gradually shifted towards the short wavelength, approaching the photoluminescence with high energy excitation. Finally, mechanism of ASPL from PbS QDs in glasses was discussed based on the phonon-assisted one-photon process.

2. Experiment

Glasses with nominal compositions of 50SiO2-25Na2O-5Al2O3-10BaO-5ZnO-2ZnS-0.8PbO (in mol%) were prepared by conventional melt-quenching method. Chemical powders with purity of >99.9% were weighted and thoroughly mixed before melting in alumina crucibles at 1400 °C for 30 min under the ambient atmosphere. The melts were quenched in a brass mold and pressed with another for quenching. The glasses thus obtained were annealed at 400 °C for 2 h to reduce the thermal stress. Annealed glasses were cut into small species for further heat treatments. PbS QDs were precipitated in the glasses by thermal treatment at temperatures near the glass transition. Average size of PbS QDs formed in the glasses was controlled by adjusting the heat-treatment temperature and duration.

Glass transition temperatures of these as-prepared glasses were found to be ~480 °C using the simultaneous thermal analyzer (STA449c/3/c/G, NETZSCH, Selb, Germany). X-ray diffraction (λ = 1.54056Å, D8 Advance, Bruker, Karlsruhe, Germany) patterns were recorded to illustrate the structural changes after thermal treatment, and the patterns were recorded at a scanning rate of 2 °/min with a step size of 0.02°. Absorption spectra of the as-prepared and heat-treated samples were recorded using a UV/Vis/NIR spectrophotometer (UV3600, Shimadzu, Kyoto, Japan). Lasers with wavelength of 800 nm, 1319 nm, and 1532 nm were used as the excitation sources. The excitation laser beams were modulated by a mechanical chopper at a frequency of 50 Hz and focused into the specimens using a silica lens with a focal length of 5 cm. Photoluminescence (PL) was collected at the direction perpendicular to the excitation beam and dispersed into a monochromator. A combination of InGaAs (or InSb) detector and lock-in amplifier was used to detect the intensity of photoluminescence.

3. Results and discussion

Figure 1 shows the X-ray diffraction patterns of as-prepared (AP) and heat-treated glasses at temperatures ranging from 530 °C and 540 °C for different durations. For the AP glass, only one broad halo was observed, indicating the non-crystalline nature of the as-prepared glass. Once the glasses were heat-treated, obvious change from yellow to dark brown and black was observed when the heat-treatment temperature and time increased from 530 °C for 8 and 10 h to 540 °C for 6 and 8 h, indicating the formation of nanocrystals in the glasses. However, due to the small size and low volume fraction of the nanocrystals, no diffraction peaks can be observed from glasses heat-treated at 530 °C for 8 h. When the heat-treated for longer time and at higher temperatures, several peaks appeared in the diffraction patterns, corresponding to the face-centered cubic PbS crystal (JCPDS No.: 65-157). These results showed that PbS nanocrystals were formed in the glasses upon thermal treatment. Average diameters of PbS nanocrystals were calculated to be 5.4 nm and 7.2 nm for glasses heat-treated at 530 °C for 10 h and 540 °C for 8 h, respectively, using the (220) diffraction peak and the Scherrer equation:

D=0.89λβcosθ
Where, D is the average size of the crystalline domains, λ is the X-ray wavelength, β is the half-width of diffraction peak in radians, and θ is the angle of diffraction.

 figure: Fig. 1

Fig. 1 X-ray diffraction patterns of as-prepared and thermal treated glasses. The bottom line is the diffraction pattern of bulk PbS crystal.

Download Full Size | PDF

Figure 2 shows the absorption and photoluminescence (PL) spectra of the heat-treated glasses. Using multiple Gaussian function simulation, it was found that the peak wavelength of the lowest absorption bands gradually shifted from 1002 nm to 1188 nm, 1323 nm and 1391 nm when the heat-treatment temperatures increased from 530 °C for 8 and 10 h to 540 °C for 6 and 8 h, respectively. Red-shift in the absorption bands with increase in the heat-treatment temperature or duration indicated the growth of the PbS QDs. Average diameters of PbS QDs were found to be 3.9 ± 0.25 nm, 4.7 ± 0.25 nm, 5.3 ± 0.27 nm, and 5.8 ± 0.34 nm for glasses heat-treated at 530 °C for 8 and 10 h, 540 °C for 6 and 8 h, respectively, using the hyperbolic band model relation [22]:

ΔE=[Eg2+22Eg(π/r)m*]1/2
Where, Eg is the bandgap energy of bulk PbS crystals, r is the radius of PbS nanocrystals and m*( = 0.085me) is the effective mass. Upon above-bandgap excitation (800 nm laser), near-infrared photoluminescence was observed. Similar to the absorption bands, peak wavelength of the photoluminescence bands shifted from 1162 nm to 1314 nm, 1402 nm, and 1465 nm when the heat-treatment temperature increased from 530 °C for 8 and 10 h to 540 °C for 6 and 8 h. Compared to the absorption bands, photoluminescence band showed obvious Stokes shift. With the size of PbS QDs increased from 3.9 nm to 4.7 nm, 5.3 nm, 5.8 nm, the Stokes shift gradually decreased from 170 meV to 101 meV, 53 meV and 45 meV, consistent with the results in our previous works which were decreased from 161 meV (3.7 nm) to 56 meV (5.2 nm) [23,24]. Such a big Stokes shift indicated that photoluminescence of PbS QDs were from the surface trap states [24].

 figure: Fig. 2

Fig. 2 Absorption and photoluminescence spectra of PbS QDs doped glasses heat-treated at different temperature and duration. The left numbers are the diameters of PbS QDs.

Download Full Size | PDF

Figure 3 shows the absorption and photoluminescence spectra of glass heat-treated at 530 °C for 10 h. Lasers with wavelength of 800 nm, 1319 nm and 1532 nm were used as the excitation. When the glass was excited at 800 nm, photoluminescence of PbS QDs appeared at 1314 nm with a Stokes shift of 101 meV. When the glass was excited at 1319 nm, where the glass had an absorption coefficient of 3.14 cm−1, one PL band with a center wavelength at 1338 nm was observed and part of the photoluminescence was located at the shorter wavelength compared to the excitation. When the glass was excited at 1532 nm, where the glass had only weak background absorption coefficient of 0.13 cm−1, the photoluminescence band was mainly composed of ASPL with a center wavelength of ~1390 nm and a blue-shift of ~82 meV compared to the excitation energy. It has to be pointed out that the ASPL even extended to ~1100 nm (1.112 eV), showing a maximum anti-Stokes shift (ΔEmax) of 318 meV. In addition, full width at half maximum (FWHM) of these photoluminescence bands increased from 107 nm to 152 nm and 177 nm when the excitation wavelength increased from 800 nm to 1319 nm and 1532 nm, respectively. It has to be pointed out that all the photoluminescence spectra observed were non-symmetric in nature.

 figure: Fig. 3

Fig. 3 Absorption spectrum (solid line), photoluminescence spectrum (open square) excited at 800 nm, 1319 nm (solid circle) and 1532 nm (open circle) of PbS QDs (diameter 4.7 nm) doped glass recorded at room temperature.

Download Full Size | PDF

Similar results are also observed in other samples in Fig. 4. Figure 4(a) shows the ASPL of 5.3 nm-sized PbS QDs upon 1532 nm laser excitation, together with the absorption spectrum and the normal PL (i.e. photoluminescence recorded upon 800 nm laser excitation) spectrum. Compared to the normal photoluminescence, the ASPL showed an obvious red-shift, indicating that the ASPL may originate a different set of defect states. With the increase in the excitation power, center wavelength of the ASPL band gradually shifted towards short wavelength, approaching the normal PL. Similar phenomenon was observed for 5.8 nm-sized PbS QDs shown in Fig. 4(b). When the excitation power increased from 0.2 W to 1.0 W, the ASPL band shifted from ~1498 nm to ~1474 nm, nearly overlapping with the normal PL band. These observations indicated that the surface trap states (which was response for the normal PL) and the defect states (which was responsible for the ASPL) were strongly dependent on the size of PbS QDs.

 figure: Fig. 4

Fig. 4 (a) Absorption spectrum (solid line), normal photoluminescence spectrum excited at 800 nm (open circle), and ASPL excited at 1532 nm (open square) with rising pump intensity of 5.3 nm-sized PbS QDs recorded at room temperature; (b) Absorption spectrum (solid line), normal photoluminescence spectrum excited at 800 nm (solid circle), and ASPL excited at 1532 nm with 0.2 W (solid square) and 1.0 W (open square) of 5.8 nm-sized PbS QDs recorded at room temperature.

Download Full Size | PDF

ASPL photoluminescence properties of 3.9 nm (excited at 1319 nm), 5.3 nm (excited at 1532 nm) and 5.8 nm (excited at 1532 nm) were shown in Fig. 5. For 3.9 nm-sized PbS QDs, photoluminescence intensity monotonically increased and peak wavelength of the ASPL band gradually blue-shifted from 1254 nm to 1240 nm as the excitation intensity increased from 50 mW to 1000 mW shown in Fig. 5(a). For 5.3 nm-sized PbS QDs, intensity of the ASPL firstly increased to the maximum as the excitation intensity increased from 100 mW to 1000 mW, and then decreased with further increase in the excitation intensity shown in Fig. 5(b). At the meantime, peak wavelength of the ASPL band gradually blue-shifted from 1462 nm to 1442 nm as the excitation intensity increased. For 5.8 nm-sized PbS QDs shown in Fig. 5(c), intensity of the ASPL also increased to the maximum as the excitation intensity increased from 10 mW to 300 mW and decreased with further increase in excitation intensity. Peak wavelength of the ASPL band gradually blue-shifted from 1498 nm to 1472 nm. The largest Stokes shifts () between the maximum ASPL energy and the excitation energy were found to be 240 meV, 395 meV and 430 meV for 3.9 nm, 5.3 nm and 5.8 nm-sized PbS QDs, respectively, much larger than most of the reported values [16,17,20].

 figure: Fig. 5

Fig. 5 ASPL photoluminescence spectra of (a) 3.9 nm (excited at 1319 nm), (b) 5.3 nm (excited at 1532 nm) and (c) 5.8 nm (excited at 1532 nm); (d) excitation power dependence of the integral ASPL intensities; (e) energies and (f) full width at half maximum of P1 and P2 bands obtained from dual Gaussian function simulation obtained from various sized PbS QDs.

Download Full Size | PDF

Integral intensity of the ASPL band was strongly dependent on the excitation power shown in Fig. 5(d). For 3.9 nm-sized PbS QDs, the integral intensity of the ASPL band showed a linear dependence on the excitation power with a slope of 0.9974. For 5.3 nm-sized PbS QDs, the integral intensity of the ASPL band initially increased linearly with a slope of 0.712 as the excitation power increased. With further increase in the excitation power, the integral intensity of the ASPL deviated from the linear and started to decrease as the increase in the excitation power. For 5.8 nm-sized PbS QDs, the integral ASPL intensity also increased linearly with a slope of 0.754 as the excitation power increased, and further increase in the excitation power led to the decrease in the integral ASPL intensity. The linear and sublinear dependence of the integral intensity of ASPL on the excitation power showed that one-photon process, instead of the two-photon excitation and Auger recombination, was responsible for the ASPL [16,20]. For 5.3 nm and 5.8 nm-sized PbS QDs, the integral intensity of ASPL band decreased with excitation power at high pump power. This phenomenon was due to the photo-induced photo-darkening, consistent with our previous work [25].

The excitation power dependent blue-shift and non-symmetry of the ASPL bands indicated that the ASPL was composed of two bands. Using dual Gaussian function simulation, one high energy band (P1) with narrow band width and one low energy band (P2) with broad band width were obtained shown in Figs. 5(e) and 5(f). It was found that the high energy bands (P1) were very close to the corresponding normal PL bands of 3.9 nm, and 5.3 nm and 5.8 nm-sized PbS QDs shown in Fig. 5(e) and Fig. 2, indicating that these P1 bands were related to the surface trap states [15,16,20]. As shown in Fig. 5(e), for 3.9 nm-sized PbS QDs, energy difference between P1 and P2 bands maintained nearly constant at 60 meV as the excitation power increased. For 5.3 nm-sized PbS QDs, energy of P1 band remained almost constant, and the energy difference between P1 and P2 bands decreased from 19.4 meV to 13.3 meV as the excitation power increased. While, for 5.8 nm-sized PbS QDs, energy of P1 band showed some variation around 0.84 eV. The energy difference between P1 and P2 bands decreased from 17.1 meV to zero as the excitation power increased to 1400 mW, and P1 and P2 bands overlapped with each other at higher excitation power. In addition, full width at half maximum (FWHM) of P1 and P2 bands showed distinct features in Fig. 5(f). For 3.9 nm-sized PbS QDs, FWHM values of P1 and P2 bands were found to be ~125 meV and ~175 meV, respectively, and FWHM values of these two bands did not change obviously with excitation power. For 5.3 nm-sized PbS QDs, P1 and P2 bands had FWHM values of ~100 meV and ~175 meV, respectively. For 5.8 nm-sized PbS QDs, FWHM value of P1 band increased from ~110 meV to ~165 meV, and that of P2 band increased from ~180 meV to ~330 meV with the increase in the excitation power. These features further confirmed that P1 and P2 bands had different origins, and probably, P2 bands were related to the defects states which located at lower energy compared to the surface trap states.

Based on the above results, the anti-Stokes photoluminescence of PbS QDs in glasses can be explained by the following model (Fig. 6). Due to the presence of dangling bonds on the surface of PbS QDs, these dangling bonds formed the surface trap states (STS) located at ΔE1 below the 1Se state of PbS QDs [15,16,24]. Besides, the defects located at the interface between the surface of PbS QDs and the glass matrix formed the defect states (DS) located at ΔE2 below the 1Se of PbS QDs. Upon above-bandgap excitation (such as 800 nm excitation used in this work), the electrons were excited from the 1Sh state into the high energy conduction levels, and these excited electrons relaxed to the 1Se states by dissipating excess energy into heat. The electrons at 1Se state can be trapped by the STS, and relaxed to the 1Sh state radiatively, yielding to normal PL (red arrow in Fig. 6). When the electron trapped at STS on the top of the 1Sh state was excited into the DS upon below-bandgap excitation (such as 1532 nm excitation, dark grey arrow), part of the electrons radiatively relaxed to the 1Sh state with the appearance of ASPL (P2 band), and part of the electrons non-radiatively relaxed to the bottom STS or 1Sh (dashed line in Fig. 6) with the dissipation of thermal energy and rise in the local temperature. With the increase in the excitation power, by absorbing the phonon energy, electrons trapped in the DS can be further excited to high energy states, even the STS, leading to the blue-shift in the ASPL, appearance of P1 band together with the P2 band and non-symmetric shape of the ASPL band. As the size of PbS QDs increased, energy gap between the STS and DS (ΔE3) decreased [26]. As a result, for large PbS QDs, energy separation between P1 and P2 decreased as the excitation power increased.

 figure: Fig. 6

Fig. 6 Schematic diagram of the photoluminescence process of PbS QDs. The red dashed lines are the surface trap states (STS) and the dashed area represents the defect states (DS).

Download Full Size | PDF

4. Conclusion

In conclusion, various size PbS QDs were precipitated in host glass by controlling temperature and time of thermal treatment. Near-infrared anti-Stokes photoluminescence (ASPL) from PbS QDs was observed upon below-bandgap excitation. Anti-Stokes shift as large as 430 meV was observed. The linear or sublinear dependence of the integral intensity of the ASPL showed the ASPL was a one-photon process. Due to the presence of defect states at the interface between PbS QDs and glass matrix, the ASPL showed strong dependence on the excitation power. The ASPL band shifted towards short wavelength side, along with the broadening the photoluminescence as the excitation power increased. The large anti-Stokes shift obtained upon below-bandgap excitation indicated that the ASPL has great potential for spectral conversion and solid state laser cooling.

Funding

Program for New Century Excellent Talents in University (No. NCET-13-0943), Chutian Scholar Program of Hubei Province, National Nature Science Foundation of China (No. 51602235).

References and links

1. R. W. Wood, “XXIX. Anti-Stokes radiation of fluorescent liquids,” London Edinburgh Dublin Philos. Magazine J. Sci. 6(35), 310–312 (1928). [CrossRef]  

2. C. Litton, D. Reynolds, T. Collins, and Y. Park, “Exciton-LO-Phonon Interaction and the Anti-Stokes Emission Line in CdS,” Phys. Rev. Lett. 25(23), 1619–1621 (1970). [CrossRef]  

3. W. Chen, A. G. Joly, and D. E. McCready, “Upconversion luminescence from CdSe nanoparticles,” J. Chem. Phys. 122(22), 224708 (2005). [CrossRef]   [PubMed]  

4. A. G. Joly, W. Chen, D. E. McCready, J.-O. Malm, and J.-O. Bovin, “Upconversion luminescence of CdTe nanoparticles,” Phys. Rev. B 71(16), 165304 (2005). [CrossRef]  

5. Z. G. Soos and R. G. Kepler, “Two-photon-absorption spectrum of poly(di-n-hexylsilane) films,” Phys. Rev. B Condens. Matter 43(14), 11908–11912 (1991). [CrossRef]   [PubMed]  

6. U. Lemmer, R. Rischer, J. Feldmann, R. Mahrt, J. Yang, A. Greiner, H. Bässler, E. Göbel, H. Heesel, and H. Kurz, “Time-resolved studies of two-photon absorption processes in poly (p-phenylenevinylene) s,” Chem. Phys. Lett. 203(1), 28–32 (1993). [CrossRef]  

7. G. Zegrya and V. Kharchenko, “New mechanism of Auger recombination of nonequilibrium current carriers in semiconductor heterostructures,” Sov. Phys. JETP 74(1), 173–181 (1992).

8. W. Heimbrodt, M. Happ, and F. Henneberger, “Giant anti-Stokes photoluminescence from semimagnetic heterostructures,” Phys. Rev. B 60(24), R16326 (1999). [CrossRef]  

9. E. Downing, L. Hesselink, J. Ralston, and R. Macfarlane, “A three-color, solid-state, three-dimensional display,” Science 273(5279), 1185–1189 (1996). [CrossRef]  

10. T. Fujino, T. Fujima, and T. Tahara, “Femtosecond fluorescence dynamics imaging using a fluorescence up-conversion microscope,” J. Phys. Chem. B 109(32), 15327–15331 (2005). [CrossRef]   [PubMed]  

11. R. Macfarlane, F. Tong, A. Silversmith, and W. Lenth, “Violet cw neodymium upconversion laser,” Appl. Phys. Lett. 52(16), 1300–1302 (1988). [CrossRef]  

12. A. Kachynski, A. Kuzmin, H. Pudavar, and P. Prasad, “Three-dimensional confocal thermal imaging using anti-Stokes luminescence,” Appl. Phys. Lett. 87(2), 023901 (2005). [CrossRef]  

13. R. I. Epstein, M. I. Buchwald, B. C. Edwards, T. R. Gosnell, and C. E. Mungan, “Observation of laser-induced fluorescent cooling of a solid,” Nature 377(6549), 500–503 (1995). [CrossRef]  

14. Y. P. Rakovich, S. Filonovich, M. Gomes, J. Donegan, D. Talapin, A. Rogach, and A. Eychmüller, “Anti-Stokes Photoluminescence in II-VI Colloidal Nanocrystals,” Phys. Status Solidi, B Basic Res. 229, 449–452 (2002). [CrossRef]  

15. E. Poles, D. C. Selmarten, O. I. Mićić, and A. J. Nozik, “Anti-Stokes photoluminescence in colloidal semiconductor quantum dots,” Appl. Phys. Lett. 75(7), 971–973 (1999). [CrossRef]  

16. X. Wang, W. W. Yu, J. Zhang, J. Aldana, X. Peng, and M. Xiao, “Photoluminescence upconversion in colloidal CdTe quantum dots,” Phys. Rev. B 68(12), 125318 (2003). [CrossRef]  

17. M. J. Fernée, P. Jensen, and H. Rubinsztein-Dunlop, “Unconventional photoluminescence upconversion from PbS quantum dots,” Appl. Phys. Lett. 91(4), 043112 (2007). [CrossRef]  

18. G. Nemova and R. Kashyap, “Laser cooling with PbSe colloidal quantum dots,” J. Opt. Soc. Am. B 29(4), 676–682 (2012). [CrossRef]  

19. J. M. Harbold and F. W. Wise, “Photoluminescence spectroscopy of PbSe nanocrystals,” Phys. Rev. B 76(12), 125304 (2007). [CrossRef]  

20. Y. P. Rakovich, S. A. Filonovich, M. J. M. Gomes, J. F. Donegan, D. V. Talapin, and A. L. Rogach, “Anti-Stokes Photoluminescence in II–VI Colloidal Nanocrystals,” Phys. Status Solidi 229, 449–452 (2002). [CrossRef]  

21. C. Liu, J. Heo, X. Zhang, and J. L. Adam, “Photoluminescence of PbS quantum dots embedded in glasses,” J. Non-Cryst. Solids 354(2), 618–623 (2008).

22. Y. Wang, A. Suna, W. Mahler, and R. Kasowski, “PbS in polymers. From molecules to bulk solids,” J. Chem. Phys. 87(12), 7315–7322 (1987). [CrossRef]  

23. N. Han, C. Liu, J. Zhang, X. Zhao, J. Heo, and Y. Jiang, “Infrared photoluminescence from lead sulfide quantum dots in glasses enriched in sulfur,” J. Non-Cryst. Solids 391, 39–42 (2014). [CrossRef]  

24. N. Han, C. Liu, Z. Zhao, J. Zhang, J. Xie, J. Han, X. Zhao, and Y. Jiang, “Quantum dots in glasses: size-dependent Stokes shift by lead chalcogenide,” Int. J. Appl. Glass Sci. 6(4), 339–344 (2015). [CrossRef]  

25. C. Liu, Y. K. Kwon, and J. Heo, “Temperature-dependent brightening and darkening of photoluminescence from PbS quantum dots in glasses,” Appl. Phys. Lett. 90(24), 241111 (2007). [CrossRef]  

26. G. Su, X. Zhao, C. Liu, and X. Zhou, “Size-dependent photoluminescence of PbS QDs embedded in silicate glasses,” J. Non-Cryst. Solids. Submitted.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 X-ray diffraction patterns of as-prepared and thermal treated glasses. The bottom line is the diffraction pattern of bulk PbS crystal.
Fig. 2
Fig. 2 Absorption and photoluminescence spectra of PbS QDs doped glasses heat-treated at different temperature and duration. The left numbers are the diameters of PbS QDs.
Fig. 3
Fig. 3 Absorption spectrum (solid line), photoluminescence spectrum (open square) excited at 800 nm, 1319 nm (solid circle) and 1532 nm (open circle) of PbS QDs (diameter 4.7 nm) doped glass recorded at room temperature.
Fig. 4
Fig. 4 (a) Absorption spectrum (solid line), normal photoluminescence spectrum excited at 800 nm (open circle), and ASPL excited at 1532 nm (open square) with rising pump intensity of 5.3 nm-sized PbS QDs recorded at room temperature; (b) Absorption spectrum (solid line), normal photoluminescence spectrum excited at 800 nm (solid circle), and ASPL excited at 1532 nm with 0.2 W (solid square) and 1.0 W (open square) of 5.8 nm-sized PbS QDs recorded at room temperature.
Fig. 5
Fig. 5 ASPL photoluminescence spectra of (a) 3.9 nm (excited at 1319 nm), (b) 5.3 nm (excited at 1532 nm) and (c) 5.8 nm (excited at 1532 nm); (d) excitation power dependence of the integral ASPL intensities; (e) energies and (f) full width at half maximum of P1 and P2 bands obtained from dual Gaussian function simulation obtained from various sized PbS QDs.
Fig. 6
Fig. 6 Schematic diagram of the photoluminescence process of PbS QDs. The red dashed lines are the surface trap states (STS) and the dashed area represents the defect states (DS).

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

D= 0.89λ βcosθ
ΔE=[ E g 2 + 2 2 E g (π/r ) m * ] 1/2
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.