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

Direct-bandgap luminescence at room-temperature from highly-strained Germanium nanocrystals

Open Access Open Access

Abstract

Efficient room-temperature luminescence at optical telecommunication wavelengths and originating from direct band-to-band recombination has been observed in tensile-strained germanium nanocrystals synthesized by mechanical grinding techniques. Selected area electron diffraction, micro-Raman and optical-absorption spectroscopy measurements indicate high tensile-strains while combined photoluminescence spectroscopy, excitation-power evolution and time-resolved measurements suggest direct band-to-band recombination. Such band-engineered germanium nanocrystals offer great possibilities for silicon-photonics integration due to their superb light-emission properties, facile fabrication and compatibility with standard microelectronic processes.

©2010 Optical Society of America

1. Introduction

Several approaches have been previously explored to provide efficient light emission in silicon and demonstrate laser action at room temperature and under electrical injection. Chemical anodization in porous silicon [1], quantum-size effects in silicon nanocrystals [2,3] or superlattices [4], Raman amplification [5,6], rare-earth elements [7,8] and point-defect centers [911] have allowed some progress towards reaching this goal. More recently, hybrid structures including III-V active layers bonded on silicon [12] and combinations of tensile strains and n-type doping in bulk germanium grown on silicon [13] have also led to most promising results.

Meanwhile, optically-active germanium (Ge) nanocrystals have been synthesized by magnetron cosputtering [1417], ion implantation [1821], isotropic chemical etching into porous germanium [22], thermal decomposition [23], solid-phase epitaxy [24], sol-gel chemistry [25] ultra-high vacuum chemical vapor deposition [26,27], Stranski-Krastanov growth by molecular-beam epitaxy [28] and, more recently, using ball-milling techniques [29]. Indeed, Ge can also provide stronger confinement than silicon nanocrystals due to a higher dielectric constant and a lighter carrier effective mass [22], while large nanocrystals can be made to emit efficiently in a spectral range more suited for optical telecommunications on chip. Moreover, band-engineered germanium nanocrystals would constitute a material of choice for silicon-photonics integration. Indeed germanium’s near-direct band structure can lead to efficient light emission at telecommunication wavelengths [13] and Ge is already known to be fully-compatible with conventional CMOS processes [30].

In this letter, we report on the efficient light-emission at room-temperature originating from direct band-to-band recombination in tensile-strained germanium nanocrystals synthesized by mechanical grinding techniques. Mechanical grinding techniques provide a convenient alternative to more conventional synthesis approaches. Indeed, we found that mechanical-grinding schemes have been sparsely used in the past to synthesize large quantities of nanocrystals made of heteroclite metal-semiconductor compounds [3135]. Moreover, some of these nanocrystalline materials have displayed unique physical properties [3537].

2. Fabrication

High-purity nanocrystalline germanium powder such shown in Fig. 1 is obtained through mechanical mortar-grinding of a bulk undoped (100) germanium wafer. A first 10 min. grinding was performed using an agate mortar and pestle to obtain a coarse powder. Then, a subsequent 25 min. grinding was performed to obtain the fine nanocrystalline powder shown in Fig. 1.

 figure: Fig. 1

Fig. 1 (a) The nanocrystalline powder synthesized by mechanical grinding and shown in the inset can also be pressed into laminates using a 12-ton press. (b,c) TEM micrographs of the Ge nanocrystals. (d) Typical selected area electron diffraction (SAED) patterns show the sound crystalline structure of the nanocrystals. The lattice constants measured from several electron diffraction patterns suggest a 2.2 ± 1.1% tensile strain compared to the bulk. (e) HRTEM micrograph showing the strain-induced deformations seen in the largest Ge nanocrystals.

Download Full Size | PDF

As shown in Fig. 1(a), this nanocrystalline powder can also be pressed into polycrystalline laminates using a 12-ton press. While mechanical grinding provides un-paralleled flexibility, low-cost and high yield, it still offers a limited level of control, especially on the grain size. As shown in Fig. 1(b,c), Ge nanocrystal diameters vary from 7 nm to hundreds of nanometers. Indeed, dynamic light scattering analysis suggests a very broad size distribution with an average nanocrystal diameter of 140 ± 20 nm and less than 10% of the nanocrystals being smaller than 20 nm, which is consistent with our SEM/TEM observations. As displayed in Fig. 1(d), typical selected area electron diffraction (SAED) patterns show the sound crystalline structure of the nanocrystals. However, the lattice constants calculated from several diffraction patterns suggest a 2.2 ± 1.1% tensile strain compared to the bulk germanium used to synthesize the powder. Indeed, HRTEM measurements performed on the largest nanocrystals and shown in Fig. 1(e) provide clear evidences of high strain-induced deformations that are consistent with previous results obtained for ball-milled nanocrystals [29].

3. Tensile-strains in bulk Germanium

Band structure calculations shown in Fig. 2(a) were obtained using an open-source software to illustrate the effect of tensile strains on the band-structure of bulk Ge [38,39]. Indeed, it illustrates how tensile strains bring the gamma (г)-point of bulk Ge closer to the L-valley minimum. For moderate tensile-strains, it has been shown that n-type doping can be used to fill the L-valley up to the gamma-point and favor direct band-to-band recombination in bulk Ge [13,40]. However, moderate levels of biaxial tensile strains can still require very high doping levels [13]. As an indicator, Fig. 2(b) shows the n-type doping concentration required to fill the L-valley as a function of the level of tensile-strains for the biaxial (2D) and hydrostatic (3D) cases calculated using a first-order band-narrowing model for bulk Ge. Yet, the results from this simple model are consistent with previous reports obtained from 0.25% biaxially tensile-strained Ge using ~7.6E + 19 cm−3 donor-impurity concentrations [13]. This model suggests that the tensile-strains measured from the SAED pattern would be sufficient to favor direct band-to-band recombination in such Ge nanocrystals.

 figure: Fig. 2

Fig. 2 (a) Electronic bandstructure calculated for fully-relaxed and 0.5% tensile-strained bulk Ge [38,39]. (b) Estimate of the donor concentration required to fill the L-valley as a function of the level of tensile-strains for the biaxial and hydrostatic cases obtained using a first-order band-narrowing model.

Download Full Size | PDF

4. Optical and structural properties

Raman and optical-absorption spectroscopy have also been used to confirm the presence of high built-in tensile strains in the nanocrystals. The clear downshift of the Ge-Ge vibrational peak shown in Fig. 3(a) is also indicator of tensile strains [4144]. Indeed, Raman width and mean roughness are both known to originate from inhomogeneous strains and dislocations [44]. While Fano resonance artifacts in the Raman spectra are most unlikely in undoped germanium [45], thermal and phonon-confinement effects were also discarded by measuring similar Ge-Ge vibrational peak’s shift and width while varying the excitation power from 1,000 to 15,000 W/cm2, as shown in Fig. 3(b). Indeed, pronounced phonon-confinement, size or thermal effects would translate into a clear linear (thermal effects) or non-linear (phonon-confinement or size-effects) variation in the Ge-Ge peaks’ shift and width when varying the excitation power over such a broad range [46,47].

 figure: Fig. 3

Fig. 3 (a) Raman spectroscopy of the Ge-Ge vibration measured for the germanium nanocrystals and the bulk material they were synthesized from. (b) Ge-Ge vibration peak shift and full-width at half-maximum (FWHM) for the nanocrystals and the bulk Ge measured at 5,000, 10,000 and 15,000 W/cm2 excitation intensities.

Download Full Size | PDF

Optical absorption spectroscopy measurements shown in Fig. 4 also indicate a much more pronounced absorption tail compared with the bulk [47], most likely originating from the downshift of the gamma-point expected from the tensile-strained germanium. Both results are consistent with the tensile strains measured from the SAED pattern.

 figure: Fig. 4

Fig. 4 Optical absorption spectroscopy measurements showing the enhanced low-energy tail for the Ge nanocrystals compared with the bulk.

Download Full Size | PDF

5. Room-temperature light-emission properties

The room-temperature photoluminescence from the nanocrystals shown in Fig. 5(a) displays a broad emission spectrum at telecommunication wavelengths, which is more than 2 orders of magnitude stronger than the bulk germanium it was synthesized from. This drastic enhancement is expected from a direct recombination-mediated regime. Most importantly, we also know that tensile strains should lower the bandgap as we have shown in Fig. 2(a). Yet, the photoluminescence spectra indicate that the quantum-confinement effect is sufficient to bring the emission from those strained nanocrystals back into the optical telecommunication range.

 figure: Fig. 5

Fig. 5 (a) Photoluminescence spectroscopy of the Ge nanocrystals (○) and bulk Ge (●) at room temperature. (b) Ge nanocrystals room-temperature integral photoluminescence intensity as a function of the excitation intensity. (c) Transient-photoluminescence measurements. We measured the detection limit of the entire system and found it better than 25 μs (highlighted region). The plain lines indicate the fits of the photoluminescence transients.

Download Full Size | PDF

The superlinear dependence of the integral photoluminescence emission as a function of the excitation power shown in Fig. 5(b) is also suggestive of direct recombination [40]. The time-resolved photoluminescence measurements obtained at room-temperature from the germanium nanocrystals and from the bulk provide more supporting evidences [48]. As shown in Fig. 5(c), the bulk germanium transients indicate that the emission possesses a fast and a slow component originating from the phonon-mediated recombination. In the Ge nanocrystals, the disappearance of the long-lived transient combined with the strongly-enhanced luminescence confirms that the emission originates mostly from direct band-to-band recombination. Although surface states might also play a role, they are highly unlikely to affect the transients measured here since their recombination lifetimes would be far below the detection limit of the setup.

6. Conclusion

In summary, we report efficient light-emission at room-temperature originating from direct band-to-band recombination in tensile-strained germanium nanocrystals synthesized by mechanical grinding techniques. Selected area electron diffraction (SAED), HRTEM, Raman & optical-absorption spectroscopy measurements point to high tensile-strains while combined photoluminescence spectroscopy, power-evolution and time-resolved measurements suggest direct band-to-band recombination. These results are also consistent with theoretical expectations for bulk germanium under high tensile strains. We believe that such band-engineered germanium nanocrystals offer great possibilities for silicon-photonics integration due to their superb room-temperature light-emission in the optical telecommunication range, facile fabrication and compatibility with standard microelectronic processes.

Acknowledgements

We are most thankful to the DARPA-MTO Young Faculty Award program and the University of Delaware Research Foundation for their support of this project.

References and links

1. A. Cullis and L. Canham, “Visible light emission due to quantum size effects in highly porous crystalline silicon,” Nature 353(6342), 335–338 (1991). [CrossRef]  

2. L. Pavesi, L. Dal Negro, C. Mazzoleni, G. Franzò, and F. Priolo, “Optical gain in silicon nanocrystals,” Nature 408(6811), 440–444 (2000). [CrossRef]   [PubMed]  

3. W. L. Wilson, P. F. Szajowski, and L. Brus, “Measurements of alloy composition and strain in thin GexSi1−x layers,” Science 262, 1242–1244 (1993). [CrossRef]   [PubMed]  

4. Z. Lu, D. J. Lockwood, and J. Baribeau, “Quantum confinement and light emission in SiO2/Si superlattices,” Nature 378(6554), 258–260 (1995). [CrossRef]  

5. H. Rong, A. Liu, R. Jones, O. Cohen, D. Hak, R. Nicolaescu, A. Fang, and M. Paniccia, “An all-silicon Raman laser,” Nature 433(7023), 292–294 (2005). [CrossRef]   [PubMed]  

6. O. Boyraz and B. Jalali, “Demonstration of a silicon Raman laser,” Opt. Express 12(21), 5269–5273 (2004), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-12-21-5269. [CrossRef]   [PubMed]  

7. H. Enner, G. Pomrenke, A. Axmann, K. Eisele, W. Haydl, and J. Schneider, “1.54-µm electroluminescence of erbium-doped silicon grown by molecular beam epitaxy,” Appl. Phys. Lett. 46(4), 381–383 (1985). [CrossRef]  

8. B. Min, T. J. Kippenberg, L. Yang, K. J. Vahala, J. Kalkman, and A. Polman, “Erbium-implanted high-Q silica toroidal microcavity laser on a silicon chip,” Phys. Rev. A 70(3), 033803 (2004). [CrossRef]  

9. S. G. Cloutier, P. A. Kossyrev, and J. M. Xu, “Optical gain and stimulated emission in periodic nanopatterned crystalline silicon,” Nat. Mater. 4(12), 887–891 (2005). [CrossRef]   [PubMed]  

10. E. Rotem, J. M. Shainline, and J. M. Xu, “Electroluminescence of nanopatterned silicon with carbon implantation and solid phase epitaxial regrowth,” Opt. Express 15(21), 14099–14106 (2007), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-15-21-14099. [CrossRef]   [PubMed]  

11. D. Recht, F. Capasso, and M. J. Aziz, “On the temperature dependence of point-defect-mediated luminescence in silicon,” Appl. Phys. Lett. 94(25), 251113 (2009). [CrossRef]  

12. A. W. Fang, H. Park, O. Cohen, R. Jones, M. J. Paniccia, and J. E. Bowers, “Electrically pumped hybrid AlGaInAs-silicon evanescent laser,” Opt. Express 14(20), 9203–9210 (2006), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-14-20-9203. [CrossRef]   [PubMed]  

13. J. Liu, X. Sun, D. Pan, X. Wang, L. C. Kimerling, T. L. Koch, and J. Michel, “Tensile-strained, n-type Ge as a gain medium for monolithic laser integration on Si,” Opt. Express 15(18), 11272–11277 (2007), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-15-18-11272. [CrossRef]   [PubMed]  

14. M. Zacharias, J. Bläsing, J. Christen, P. Veit, B. Dietrich, and D. Bimberg, “Formation of Ge nanocrystals with sharp size distribution: structural and optical characterization,” Superlattices Microstruct. 18(2), 139–146 (1995). [CrossRef]  

15. Y. Maeda, N. Tsukamoto, Y. Yazawa, Y. Kanemitsu, and Y. Masumoto, “Visible photoluminescence of Ge microcrystals embedded in SiO2 glassy matrices,” Appl. Phys. Lett. 59(24), 3168–3170 (1991). [CrossRef]  

16. W. K. Choi, V. Ng, S. P. Ng, H. H. Thio, Z. X. Shen, and W. S. Li, “Raman characterization of germanium nanocrystals in amorphous silicon oxide films synthesized by rapid thermal annealing,” J. Appl. Phys. 86(3), 1398–1403 (1999). [CrossRef]  

17. S. Takeoka, M. Fujii, S. Hayashi, and K. Yamamoto, “Size-dependent near-infrared photoluminescence from Ge nanocrystals embedded in SiO2 matrices,” Phys. Rev. B 58(12), 7921–7925 (1998). [CrossRef]  

18. Y. Kanemitsu, K. Masuda, M. Yamamoto, K. Kajiyama, and T. Kushida, “Near-infrared photoluminescence from Ge nanocrystals in SiO2 matrices,” J. Lumin. 87–89, 457–459 (2000). [CrossRef]  

19. Y. Kanemitsu, H. Uto, Y. Masumoto, and Y. Maeda, “On the origin of visible photoluminescence in nanometer-size Ge crystallites,” Appl. Phys. Lett. 61(18), 2187–2190 (1992). [CrossRef]  

20. D. C. Paine, C. Caragianis, T. Y. Kim, Y. Shigesato, and T. Ishahara, “Visible photoluminescence from nanocrystalline Ge formed by H2 reduction of Si0.6Ge0.4O2,” Appl. Phys. Lett. 62(22), 2842–2844 (1993). [CrossRef]  

21. S. Okamoto and Y. Kanemitsu, “Photoluminescence properties of surface-oxidized Ge nanocrystals: Surface localization of excitons,” Phys. Rev. B 54(23), 16421–16424 (1996). [CrossRef]  

22. G. Kartopu, A. V. Sapelkin, V. A. Karavanskii, U. Serincan, and R. Turan, “Structural and optical properties of porous nanocrystalline Ge,” J. Appl. Phys. 103(11), 113518 (2008). [CrossRef]  

23. H. P. Wu, M. Y. Ge, C. W. Cao, Y. W. Wang, Y. W. Zeng, L. N. Wang, G. Q. Zhang, and J. Z. Jiang, “Blue emission of Ge nanocrystals prepared by thermal decomposition,” Nanotechnology 17(21), 5339–5343 (2006). [CrossRef]  

24. Q. Xiao and H. Tu, “Ge–Si system nanoclusters in Si matrix formed by solid-phase epitaxy,” Appl. Phys. Lett. 86(20), 201914 (2005). [CrossRef]  

25. H. Yang, X. Wang, H. Shi, S. Xie, F. Wang, X. Gu, and X. Yao, “Photoluminescence of Ge nanoparticles embedded in SiO2 glasses fabricated by a sol–gel method,” Appl. Phys. Lett. 81(27), 5144–5146 (2002). [CrossRef]  

26. E. Palange, G. Capelli, L. Di Gaspare, and F. Evangelisti, “Atomic force microscopy and photoluminescence study of Ge layers and self-organized Ge quantum dots on Si(100),” Appl. Phys. Lett. 68(21), 2982–2984 (1996). [CrossRef]  

27. S. W. Lee, P. S. Chen, S. L. Cheng, M. H. Lee, H. T. Chang, C.-H. Lee, and C. W. Liu, “Modified growth of Ge quantum dots using C2H4 mediation by ultra-high vacuum chemical vapor deposition,” Appl. Surf. Sci. 254(19), 6261–6264 (2008). [CrossRef]  

28. N. Sustersic, L. Nataraj, C. Weiland, M. Coppinger, M. V. Shaleev, A. V. Novikov, R. Opila, S. G. Cloutier, and J. Kolodzey, “High power N-face GaN high electron mobility transistors grown by molecular beam epitaxy with optimization of AlN nucleation,” Appl. Phys. Lett. 94, 183103 (2009). [CrossRef]  

29. P. K. Giri, “Strain analysis on freestanding germanium nanocrystals,” J. Phys. D Appl. Phys. 42(24), 245402 (2009). [CrossRef]  

30. D. J. Friedman, M. Meghelli, B. D. Parker, J. Yang, H. A. Ainspan, A. V. Rylyakov, Y. H. Kwark, M. B. Ritter, L. Shan, S. J. Zier, M. Sorna, and M. Soyuer, “SiGe BiCMOS integrated circuits for high-speed serial communication links,” IBM J. Res. Develop. 47, 259–282 (2003). [CrossRef]  

31. P. Marin, A. Hernando, M. Lopez, T. Kulik, L. K. Varga, and G. Hadjipanayis, “Influence of mechanical grinding on the structure and magnetic properties of FeCuNbSiB material,” J. Magn. Magn. Mater. 272–276, E1131–E1133 (2004). [CrossRef]  

32. M. Lopez, P. Marin, P. Agudo, I. Carabias, J. de la Venta, and A. Hernando, “Nanocrystalline FeSiBNbCu alloys: Differences between mechanical and thermal crystallization process in amorphous precursors,” J. Alloy. Comp. 434–435, 199–202 (2007). [CrossRef]  

33. G. F. Goya, “Nanocrystalline CuFe2O4 obtained by mechanical grinding,” J. Mater. Sci. Lett. 16(7), 563–565 (1997). [CrossRef]  

34. C. C. Koch and Y. S. Cho, “Nanocrystals by high energy ball milling,” Nanostructured Materials 1(3), 207–212 (1992). [CrossRef]  

35. V. Svrcek, J.-L. Rehspringer, E. Gaffet, A. Slaoui, and J.-C. Muller, “Unaggregated silicon nanocrystals obtained by ball milling,” J. Cryst. Growth 275(3-4), 589–597 (2005). [CrossRef]  

36. Q. Li, C. Liu, Z. Liu, and Q. Gong, “Broadband optical limiting and two-photon absorption properties of colloidal GaAs nanocrystals,” Opt. Express 13(6), 1833–1838 (2005), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-13-6-1833. [CrossRef]   [PubMed]  

37. G. L. Tan and X. F. Yu, “Capping the Ball-Milled CdSe Nanocrystals for Light Excitation,” J. Phys. Chem. C 113(20), 8724–8729 (2009). [CrossRef]  

38. A. Paul, M. Luisier, R. Kim, M. McLennan, M. Lundstrom, and G. Klimeck, “Band Structure Lab” (2006), http://hdl.handle.net/ 10254/nanohub-r1308.6.

39. G. Klimeck, M. Korkusinski, F. Saied, H. Xu, S. Lee, M. Sayeed, and S. Goasguen, “Development of a Nanoelectronic 3-D (NEMO 3-D) Simulator for Multimillion Atom Simulations and Its Application to Alloyed Quantum Dots,” Comput. Model. Eng. Sci. 3, 601–642 (2002).

40. X. Sun, J. Liu, L. C. Kimerling, and J. Michel, “Room-temperature direct bandgap electroluminesence from Ge-on-Si light-emitting diodes,” Opt. Lett. 34(8), 1198–1200 (2009). [CrossRef]   [PubMed]  

41. J. C. Tsang, P. M. Mooney, F. Dacol, and J. O. Chu, “Measurements of alloy composition and strain in thin GexSi1−x layers,” J. Appl. Phys. 75(12), 8098–8108 (1994). [CrossRef]  

42. I. De Wolf, “Micro-Raman spectroscopy to study local mechanical stress in silicon integrated circuits,” Semicond. Sci. Technol. 11(2), 139–154 (1996). [CrossRef]  

43. T. Tezuka, N. Sugiyama, and S. Takagi, “Fabrication of strained Si on an ultrathin SiGe-on-insulator virtual substrate with a high-Ge fraction,” Appl. Phys. Lett. 79(12), 1798–1800 (2001). [CrossRef]  

44. J. Camassel, L. A. Falkovsky, and N. Planes, “Photoluminescence properties of surface-oxidized Ge nanocrystals: Surface localization of excitons,” Phys. Rev. B 63, 035309 (2000). [CrossRef]  

45. N. H. Nickel, P. Lengsfdeld, and I. Sieber, “Raman spectroscopy of heavily doped polycrystalline silicon thin films,” Phys. Rev. B 61(23), 15558–15561 (2000). [CrossRef]  

46. S. G. Cloutier, R. S. Guico, and J. M. Xu, “Phonon localization in periodic uniaxially nanostructured silicon,” Appl. Phys. Lett. 87(22), 222104 (2005). [CrossRef]  

47. S. G. Cloutier, C.-H. Hsu, P. A. Kossyrev, and J. Xu, “Radiative recombination enhancement in silicon via phonon localization and selection-rule breaking,” Adv. Mater. 18, 841–844 (2006). [CrossRef]  

48. B. V. Kamenev, L. Tsybeskov, J.-M. Baribeau, and D. J. Lockwood, “Coexistence of fast and slow luminescence in three-dimensional Si/Si1−xGex nanostructures,” Phys. Rev. B 72(19), 193306 (2005). [CrossRef]  

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 (5)

Fig. 1
Fig. 1 (a) The nanocrystalline powder synthesized by mechanical grinding and shown in the inset can also be pressed into laminates using a 12-ton press. (b,c) TEM micrographs of the Ge nanocrystals. (d) Typical selected area electron diffraction (SAED) patterns show the sound crystalline structure of the nanocrystals. The lattice constants measured from several electron diffraction patterns suggest a 2.2 ± 1.1% tensile strain compared to the bulk. (e) HRTEM micrograph showing the strain-induced deformations seen in the largest Ge nanocrystals.
Fig. 2
Fig. 2 (a) Electronic bandstructure calculated for fully-relaxed and 0.5% tensile-strained bulk Ge [38,39]. (b) Estimate of the donor concentration required to fill the L-valley as a function of the level of tensile-strains for the biaxial and hydrostatic cases obtained using a first-order band-narrowing model.
Fig. 3
Fig. 3 (a) Raman spectroscopy of the Ge-Ge vibration measured for the germanium nanocrystals and the bulk material they were synthesized from. (b) Ge-Ge vibration peak shift and full-width at half-maximum (FWHM) for the nanocrystals and the bulk Ge measured at 5,000, 10,000 and 15,000 W/cm2 excitation intensities.
Fig. 4
Fig. 4 Optical absorption spectroscopy measurements showing the enhanced low-energy tail for the Ge nanocrystals compared with the bulk.
Fig. 5
Fig. 5 (a) Photoluminescence spectroscopy of the Ge nanocrystals (○) and bulk Ge (●) at room temperature. (b) Ge nanocrystals room-temperature integral photoluminescence intensity as a function of the excitation intensity. (c) Transient-photoluminescence measurements. We measured the detection limit of the entire system and found it better than 25 μs (highlighted region). The plain lines indicate the fits of the photoluminescence transients.
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.