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

Dynamically tuning emission band of CdSe/ZnS quantum dots assembled on Ag nanorod array: plasmon-enhanced Stark shift

Open Access Open Access

Abstract

We demonstrate tuning emission band of CdSe/ZnS semiconductor quantum dots (SQDs) closely-packed in the proximity of Ag nanorod array by dynamically adjusting exciton-plasmon interaction. Large red-shift is observed in two-photon luminescence (TPL) spectra of the SQDs when the longitudinal surface plasmon resonance (LSPR) of Ag nanorod array is adjusted to close to excitation laser wavelength, and the spectral red-shift of TPL reaches as large as 101 meV by increasing excitation power, which is slightly larger than full width at half-maximum of emission spectrum of the SQDs. The observed LSPR-dependent spectral shifting behaviors are explained by a theoretical model of plasmon-enhanced quantum-confined Stark effect. These observations could find the applications in dynamical information processing in active plasmonic and photonic nanodevices.

©2011 Optical Society of America

1. Introduction

Tuning absorption and emission band of optical nanoemitters has prospective applications in optical information processing, such as optical modulating and switching [1]. Semiconductor quantum dots with tunable energy levels are the solid state “artificial atoms” and the building blocks for quantum information [24]. Strong surface plasmon resonance (SPR) of metallic nanostructures largely enhances the local field at the excitation and/or emission frequency of optical nanoemitters [57], which offers a powerful tool to tune optical properties of optical nanoemitters [811]. SPR could effectively modulate emission spectrum of multilevel nanoemitters by enhancing resonant transitions and suppressing nonresonant transitions [12,13]. Moreover, near-field coherent coupling between plasmon and exciton forms plexciton and generates hybridized absorption band [1417].

Quantum-confined Stark effect in low-dimensional semiconductor materials due to strong confinement of electron and hole in a bound exciton leads to a prominent power-dependent absorption band shift [1827]. Plasmon-enhanced Stark shifts in Au/CdSe core/shell nanoparticles have been demonstrated by Zhang et al. [28]. On the one hand, by tuning gap distance of gold nanoparticle dimer to selectively enhance spontaneous emission rate of the favor transitions, shaping emission spectra of multilevel fluorescent molecules is demonstrated by Ringler et al. [12]. On the other hand, by tuning nanoparticle size to overlap the favor Raman and plasmon modes, plasmonic control of the Raman spectrum of a single molecule in a gold-silver core-shell nanoparticle dimer is reported by Dadosh et al. [13]. Interestingly, spectral narrowing and red-shifting of CdSe/ZnS semiconductor quantum dots (SQDs) on the Ag island film is observed by Soganci et al. [29], but the detailed mechanism of spectral narrowing and shifting is unclear.

In this paper, we investigate power-dependent red-shift of two-photon luminescence (TPL) of CdSe/ZnS SQDs in the proximity of Ag nanorods (AgNRs) array, optimize spectral red-shift by tuning longitudinal surface plasmon resonances of the AgNRs array, then discuss the underlying physics processes, i.e. plasmon-enhanced quantum-confined Stark effect.

2. Synthesis and characterization

Our arrayed AgNRs were grown by using anodic aluminum oxide (AAO). AAO templates were fabricated by using a two-step anodization process [30]. The anodization voltage was stepwise reduced from 19 V down to 6 V to decrease the thickness of Al2O3 barrier layer. Silver nanorods were deposited in the pores of AAO templates by alternating current electrolysis (50 Hz, 5 V ac) in an electrolyte containing AgNO3 (0.03 M) and H2SO4 acid (0.03 M) with a Pt counter electrode. The CdSe/ZnS core-shell SQDs were purchased from Invitrogen Corporation and drop-coated onto the barrier layer of AAO template. The diameter of SQDs is 13.4 ± 0.7 nm [31].

The AAO templates loaded with AgNRs were etched by a precise Ar ion polishing system (Gatan PIPS Model 691) and examined by scanning electron microscopy (SEM). The SEM was performed by using a FEG SEM Sirion 200 operated at an accelerating voltage of 25.0 kV. The TEM was performed by using a JEOL 2010HT operated at 100 kV. The absorption spectra were recorded by a UV-VIS-NIR spectrophotometer (Varian Cary 5000) by using a p-polarized source with an incident angle of approximate 80°. The excitation source for the measurements of SQDs TPL was generated by a mode lock Ti:sapphire laser (Mira 900, Coherent) with a pulse width approximate 3 ps and a repetition rate of 76 MHz, and the excitation wavelength was set to 804 nm. The laser scattering noise was blocked by a band pass filter. A lens with 70 mm focal-length was used to focus laser beam onto the sample. The incident angle of laser beam was approximately 80° and the area of focus spot on the sample was about 1.32 × 10−3 mm2. The TPL from the sample was collected by reflection measurement, filtered by a couple filters and recorded by spectrometry (Spectrapro 2500i, Acton) with a liquid-nitrogen-cooled CCD (SPEC-10, Princeton). All the TPL measurements in the report were carried out at room temperature (~25°C).

3. Results and discussion

Figure 1 presents the nanostructures and absorption spectra of samples. Figure 1(a) illustrates the side view sketch of the AgNR array and CdSe/ZnS quantum dots nanocomplex. The thickness of alumina barrier layer (Δ) is about 7 nm. Figure 1(b) shows the SEM image of the AAO template loaded with AgNRs at the barrier side, with the thin Al2O3 barrier layer being removed by a precise Ar ion polishing system (Gatan PIPS Model 691). The diameter dAg and period of the AgNRs is about 18 ± 3 nm and 45 ± 4 nm. The TEM image of the AgNRs with λLSPR = 700 nm is shown in Fig. 1(c), and the length lAg is measured to be about 90 nm. The length of AgNRs is adjusted conveniently by varying the growth time tg, and as a consequence the different LSPR peaks are obtained in the AgNR array nanosystem [30]. Figure 1(d) presents the absorption spectra of the AgNR arrays with λLSPR = 650, 670, 700, 710, 755 and 795 nm. The recorded incident angle (θin) is approximately 80°, and the oscillations in the absorption spectra are attributed to the surface interferences of the AAO templates.

 figure: Fig. 1

Fig. 1 Nanostructures and absorption spectra of samples. (a) Schematic of SQD-AgNR nanocomplex. (b) SEM image of the AAO template loaded with AgNRs at the barrier layer side. (c) TEM image of the AgNRs with λLSPR = 700 nm. (d) Absorption spectra (θin = 80°) of AgNR arrays with λLSPR = 650, 670, 700, 710, 755 and 795 nm.

Download Full Size | PDF

Figure 2 shows normalized TPL spectra of CdSe/ZnS SQDs coated on the Ag nanorod array at excitation power Pexc = 5, 15, 20, 25, 30, 35, 50 and 65 mW. The LSPR wavelength λLSPR of the AgNRs is 670 nm, and the excitation wavelength λlaser is 804 nm. The excitation laser is p-polarized and the electric field along the long axis of AgNR array is E|| = E0 sinθin / nAAO, where θin is the incident angle and nAAO is the refractive index of AAO template. From Fig. 2, one can see that the TPL peak red-shifts as the excitation power increases. The total re-shift reaches about 33 nm = 88.4 meV when Pexc increases from 5 mW to 65 mW, which is slightly larger than the full width at half-maximum (FWHM) of the emission spectrum of the CdSe/ZnS SQDs. The PL spectra of pure AgNRs mainly shows scattering laser light under the used excitation power. A typical shift rate of less than 0.1 nm/K was observed in the temperature-dependence PL spectra of the SQDs in the range of 5-80 °C. This result accords with previous report [32], and indicates the thermal effect is not the dominant physical process for our observed red-shifted luminescence.

 figure: Fig. 2

Fig. 2 TPL spectra of SQD-AgNR nanocomplex with λLSPR = 670 nm and Pexc = 5, 15, 20, 25, 30, 35, 50 and 65 mW. The emission peak shifts from ~665 nm to ~698 nm as the excitation power increases from 5 mW to 65 mW. The edge ~740 nm in TPL spectra is caused by the band-pass filter.

Download Full Size | PDF

Figure 3 gives the normalized TPL spectra of the SQD-AgNR nanocomplex recorded at a fixed excitation power 65 mW. The LSPR wavelengths of the AgNR arrays are 555, 577, 610, 650, 670, 700, 710, 755, and 795 nm, respectively. It clearly shows that the TPL peak is tuned from 665 nm to 710 nm by adjusting LSPR of the AgNRs array from 555 nm to 795 nm with a fixed excitation power.

 figure: Fig. 3

Fig. 3 TPL spectra of SQD-AgNR nanocomplex with λLSPR = 555, 577, 610, 650, 670, 700, 710, 755, and 795 nm, the excitation power is fixed at 65 mW.

Download Full Size | PDF

Power-dependent TPL peak wavelength of the SQD-AgNR nanocomplex are plotted in Fig. 4 , which clearly shows that the spectral redshift of SQD-AgNR nanocomplex is approximately proportional to the excitation power Pexc when Pexc is not too strong. The nanocomplex with shorter AgNRs (smaller λLSPR and far off-resonance to laser wavelength λlaser) has a smaller spectral shift. The spectral shift becomes saturated at strong excitation region (Pexc > 35 mW) in the nanocomplex with λLSPR = 795 nm, which is very close to excitation laser wavelength (λlaser = 804 nm), and the emission peak changes from 664 nm (1.867 eV) to 702 nm (1.764 eV) with a shift as large as about 38 nm (101 meV). Also, it is observed that the emission peak of SQDs on AAO template without AgNRs almost remains stable (shifts about 1.6 nm).The observed spectral shift in the weak excitation region are approximately summarized by the relationship

ΔλΔλ0+RShiftPexc
where RShift is the rate of spectral shifting, which becomes larger when λLSPR becomes close to λlaser. Δλ0 is a power-independent spectral shift caused by plasmon-mediated Förster resonant energy transfer (FRET) betweem the closely packed SQDs with a size distribution [33,34]. Δλ0 of SQDs assembly in the absence of AgNRs array is supposed to be zero. Note that RShift of the controlled SQDs sample is only about 0.024 nm/mW, while RShift of the SQD-AgNR nanocomplex with λLSPR = 795 nm reaches as faster as 0.81 nm/mW. It means that spectral shifting rate of CdSe/ZnS SQDs is enhanced about 34 times by the AgNRs array.

 figure: Fig. 4

Fig. 4 Peak emission wavelength of the SQD-AgNR nanocomplex as a function of excitation power. Here, LSPR wavelengths of the AgNR arrays are 577, 610, 650, 670, 700 and 795 nm.

Download Full Size | PDF

We turn to theoretical analysis of the underlying physics mechanism of LSPR-dependent spectral shift in our SQD-AgNR nanocomplex. The Stark shift of spectral lines is enhanced by bound excitons in semiconductor nanostructures due to quantum confinement. The electron and hole confined in bound excitons are pulled in the opposite direction by the applied electric field. Strong light-matter coupling could also leads to Stark shift, and the amount of shift ΔEStark is proportional to the local light intensity |Elocal|2 [18,28,35]. Plasmon resonance significantly enhances quantum-confined Stark effect due to strong enhancement of local field. Moreover, strong plasmon-exciton coupling may result in boexcitonic interaction which can induce the spectral redshift [3639]; the formula of Stark shift is modified as

ΔEStarkPexc(ωLaserωLSPR)2+γ2

where 1/[(ωLaserωLSPR)2+γ2] represents the dispersion behavior of the local field enhancement factor |f|2, γ is a damping factor, ωLaser and ωLSPR are the optical frequency of laser pulses and LSPR of the AgNR array, respectively. Equation (2) well explains observed behaviors of the spectral shift: (i) ΔλStark (= ΔλStark – Δλ0) is proportional to excitation power Pexc; and (ii) ΔλStark saturates when ωLSPR is close to ωLaser. The spectral shifting rate can be calculated from the relation

RShift=ΔEStarkPexc=A(ωLaserωLSPR)2+γ2

The experimental data and theoretical fitting curve of RShift as a function of λLSPR of AgNRs array are plotted in Fig. 5 , in which, the fitting parameters γ = 0.445 meV. One can see that the theoretical fitting coincide with the experimental data very well. Based on these experimental observations and theoretical analysis, we think that the giant spectral red-shift in SQD-AgNR nanocomplex is mainly attributed to the plasmon-enhanced quantum-confined effect.

 figure: Fig. 5

Fig. 5 Spectral shifting rate RShift of the SQD-AgNR nanocomplex as a function of LSPR wavelength of the AgNRs array.

Download Full Size | PDF

4. Conclusion

In summary, we found giant dynamically red-shifting of TPL of CdSe/ZnS SQDs closely-packed in the proximity of Ag nanorods array. By tuning LSPR of AgNR arrays to close to the excitation laser wavelength, the spectral red-shift reaches approximately 101 meV, which is slightly larger than the FWHM of the emission spectrum of the CdSe/ZnS SQDs. LSPR-dependent shifting rate RShift of the SQD-AgNR nanocomposite is well explained by plasmon-enhanced quantum-confined Stark effect. These observations could find the applications in plasmon-based information processing nanodevices, such as optical modulating and switching.

Acknowledgments

This work was supported in part by NSFC (10874134, 10874020), National Basic Research Program of China (2011CB922200, 2010CB923200), Key Project of Ministry of Education of China (708063), and the Fundamental Research Funds for the Central Universities (20102020101000025).

References and links

1. D. Pacifici, H. J. Lezec, and H. A. Atwater, “All-optical modulation by plasmonic excitation of CdSe quantum dots,” Nat. Photonics 1(7), 402–406 (2007). [CrossRef]  

2. L. Bányai and S. W. Koch, Semiconductor Quantum Dots (World Scientific Publishing Co. Pte. Ltd., 1993).

3. Q. Q. Wang, A. Muller, M. T. Cheng, H. J. Zhou, P. Bianucci, and C. K. Shih, “Coherent control of a V-type three-level system in a single quantum dot,” Phys. Rev. Lett. 95(18), 187404 (2005). [CrossRef]   [PubMed]  

4. A. Gonzalez-Tudela, D. Martin-Cano, E. Moreno, L. Martin-Moreno, C. Tejedor, and F. J. Garcia-Vidal, “Entanglement of two qubits mediated by one-dimensional plasmonic waveguides,” Phys. Rev. Lett. 106(2), 020501 (2011). [CrossRef]   [PubMed]  

5. H. Xu, E. J. Bjerneld, M. Käll, and L. Börjesson, “Spectroscopy of single hemoglobin molecules by surface enhanced raman scattering,” Phys. Rev. Lett. 83(21), 4357–4360 (1999). [CrossRef]  

6. S. Kühn, U. Håkanson, L. Rogobete, and V. Sandoghdar, “Enhancement of single-molecule fluorescence using a gold nanoparticle as an optical nanoantenna,” Phys. Rev. Lett. 97(1), 017402 (2006). [CrossRef]   [PubMed]  

7. A. Kinkhabwala, Z. Yu, S. Fan, Y. Avlasevich, K. Müllen, and W. E. Moerner, “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nat. Photonics 3(11), 654–657 (2009). [CrossRef]  

8. H. A. Atwater, “The promise of plasmonics,” Sci. Am. 296(4), 56–62 (2007). [CrossRef]   [PubMed]  

9. P. Mühlschlegel, H.-J. Eisler, O. J. F. Martin, B. Hecht, and D. W. Pohl, “Resonant optical antennas,” Science 308(5728), 1607–1609 (2005). [CrossRef]   [PubMed]  

10. Z. C. Dong, X. L. Zhang, H. Y. Gao, Y. Luo, C. Zhang, L. G. Chen, R. Zhang, X. Tao, Y. Zhang, J. L. Yang, and J. G. Hou, “Generation of molecular hot electroluminescence by resonant nanocavity plasmons,” Nat. Photonics 4(1), 50–54 (2010). [CrossRef]  

11. N. Liu, M. L. Tang, M. Hentschel, H. Giessen, and A. P. Alivisatos, “Nanoantenna-enhanced gas sensing in a single tailored nanofocus,” Nat. Mater. 10(8), 631–636 (2011). [CrossRef]   [PubMed]  

12. M. Ringler, A. Schwemer, M. Wunderlich, A. Nichtl, K. Kürzinger, T. A. Klar, and J. Feldmann, “Shaping emission spectra of fluorescent molecules with single plasmonic nanoresonators,” Phys. Rev. Lett. 100(20), 203002 (2008). [CrossRef]   [PubMed]  

13. T. Dadosh, J. Sperling, G. W. Bryant, R. Breslow, T. Shegai, M. Dyshel, G. Haran, and I. Bar-Joseph, “Plasmonic control of the shape of the Raman spectrum of a single molecule in a silver nanoparticle dimer,” ACS Nano 3(7), 1988–1994 (2009). [CrossRef]   [PubMed]  

14. N. T. Fofang, T. H. Park, O. Neumann, N. A. Mirin, P. Nordlander, and N. J. Halas, “Plexcitonic nanoparticles: plasmon-exciton coupling in nanoshell-J-aggregate complexes,” Nano Lett. 8(10), 3481–3487 (2008). [CrossRef]   [PubMed]  

15. A. Manjavacas, F. J. García de Abajo, and P. Nordlander, “Quantum plexcitonics: strongly interacting plasmons and excitons,” Nano Lett. 11(6), 2318–2323 (2011). [CrossRef]   [PubMed]  

16. R. D. Artuso and G. W. Bryant, “Optical response of strongly coupled quantum dot-metal nanoparticle systems: double peaked Fano structure and bistability,” Nano Lett. 8(7), 2106–2111 (2008). [CrossRef]   [PubMed]  

17. N. I. Cade, T. Ritman-Meer, and D. Richards, “Strong coupling of localized plasmons and molecular excitons in nanostructured silver films,” Phys. Rev. B 79(24), 241404 (2009). [CrossRef]  

18. S. A. Empedocles and M. G. Bawendi, “Quantum-confined stark effect in single CdSe nanocrystallite quantum dots,” Science 278(5346), 2114–2117 (1997). [CrossRef]   [PubMed]  

19. M. E. Flatté, A. A. Kornyshev, and M. Urbakh, “Giant Stark effect in quantum dots at liquid/liquid interfaces: a new option for tunable optical filters,” Proc. Natl. Acad. Sci. U.S.A. 105(47), 18212–18214 (2008). [CrossRef]   [PubMed]  

20. M. Joffre, D. Hulin, A. Migus, and M. Combescot, “Laser-induced exciton splitting,” Phys. Rev. Lett. 62(1), 74–77 (1989). [CrossRef]   [PubMed]  

21. A. Muller, W. Fang, J. Lawall, and G. S. Solomon, “Creating polarization-entangled photon pairs from a semiconductor quantum dot using the optical Stark effect,” Phys. Rev. Lett. 103(21), 217402 (2009). [CrossRef]   [PubMed]  

22. X. Xu, B. Sun, E. D. Kim, K. Smirl, P. R. Berman, D. G. Steel, A. S. Bracker, D. Gammon, and L. J. Sham, “Single charged quantum dot in a strong optical field: absorption, gain, and the ac-Stark effect,” Phys. Rev. Lett. 101(22), 227401 (2008). [CrossRef]   [PubMed]  

23. C. Sieh, T. Meier, F. Jahnke, A. Knorr, S. W. Koch, P. Brick, M. Hübner, C. Ell, J. Prineas, G. Khitrova, and H. M. Gibbs, “Coulomb memory signatures in the excitonic optical Stark effect,” Phys. Rev. Lett. 82(15), 3112–3115 (1999). [CrossRef]  

24. K. C. Je, H. Ju, M. Treguer, T. Cardinal, and S. H. Park, “Local field-induced optical properties of Ag-coated CdS quantum dots,” Opt. Express 14(17), 7994–8000 (2006). [CrossRef]   [PubMed]  

25. G. W. Wen, J. Y. Lin, H. X. Jiang, and Z. Chen, “Quantum-confined Stark effects in semiconductor quantum dots,” Phys. Rev. B Condens. Matter 52(8), 5913–5922 (1995). [CrossRef]   [PubMed]  

26. T. Unold, K. Mueller, C. Lienau, T. Elsaesser, and A. D. Wieck, “Optical Stark effect in a quantum dot: ultrafast control of single exciton polarizations,” Phys. Rev. Lett. 92(15), 157401 (2004). [CrossRef]   [PubMed]  

27. B. J. Sussman, J. G. Underwood, R. Lausten, M. Y. Ivanov, and A. Stolow, “Quantum control via the dynamic Stark effect: Application to switched rotational wave packets and molecular axis alignment,” Phys. Rev. A 73(5), 053403 (2006). [CrossRef]  

28. J. Zhang, Y. Tang, K. Lee, and M. Ouyang, “Tailoring light-matter-spin interactions in colloidal hetero-nanostructures,” Nature 466(7302), 91–95 (2010). [CrossRef]   [PubMed]  

29. I. M. Soganci, S. Nizamoglu, E. Mutlugun, O. Akin, and H. V. Demir, “Localized plasmon-engineered spontaneous emission of CdSe/ZnS nanocrystals closely-packed in the proximity of Ag nanoisland films for controlling emission linewidth, peak, and intensity,” Opt. Express 15(22), 14289–14298 (2007). [CrossRef]   [PubMed]  

30. Z. K. Zhou, M. Li, Z. J. Yang, X. N. Peng, X. R. Su, Z. S. Zhang, J. B. Li, N. C. Kim, X. F. Yu, L. Zhou, Z. H. Hao, and Q. Q. Wang, “Plasmon-mediated radiative energy transfer across a silver nanowire array via resonant transmission and subwavelength imaging,” ACS Nano 4(9), 5003–5010 (2010). [CrossRef]   [PubMed]  

31. H. Kobayashi, Y. Hama, Y. Koyama, T. Barrett, C. A. S. Regino, Y. Urano, and P. L. Choyke, “Simultaneous multicolor imaging of five different lymphatic basins using quantum dots,” Nano Lett. 7(6), 1711–1716 (2007). [CrossRef]   [PubMed]  

32. G. W. Walker, V. C. Sundar, C. M. Rudzinski, A. W. Wun, M. G. Bawendi, and D. G. Nocera, “Quantum-dot optical temperature probes,” Appl. Phys. Lett. 83(17), 3555–3557 (2003). [CrossRef]  

33. A. O. Govorov, J. Lee, and N. A. Kotov, “Theory of plasmon-enhanced Förster energy transfer in optically excited semiconductor and metal nanoparticle,” Phys. Rev. B 76(12), 125308 (2007). [CrossRef]  

34. M. Durach, A. Rusina, V. I. Klimov, and M. I. Stockman, “Nanoplasmonic renormalization and enhancement of Coulomb interactions,” New J. Phys. 10(10), 105011 (2008). [CrossRef]  

35. A. Kaplan, M. F. Andersen, and N. Davidson, “Suppression of inhomogeneous broadening in rf spectroscopy of optically trapped atoms,” Phys. Rev. A 66(4), 045401 (2002). [CrossRef]  

36. M. A. Mahmoud, A. J. Poncheri, R. L. Phillips, and M. A. El-Sayed, “Plasmonic field enhancement of the exciton-exciton annihilation process in a poly(p-phenyleneethynylene) fluorescent polymer by Ag nanocubes,” J. Am. Chem. Soc. 132(8), 2633–2641 (2010). [CrossRef]   [PubMed]  

37. H. Aouani, S. Itzhakov, D. Gachet, E. Devaux, T. W. Ebbesen, H. Rigneault, D. Oron, and J. Wenger, “Colloidal quantum dots as probes of excitation field enhancement in photonic antennas,” ACS Nano 4(8), 4571–4578 (2010). [CrossRef]   [PubMed]  

38. M. Combescot and R. Combescot, “Optical Stark effect of the exciton: biexcitonic origin of the shift,” Phys. Rev. B Condens. Matter 40(6), 3788–3801 (1989). [CrossRef]   [PubMed]  

39. D. Hulin and M. Joffre, “Excitonic optical Stark redshift: the biexciton signature,” Phys. Rev. Lett. 65(27), 3425–3428 (1990). [CrossRef]   [PubMed]  

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 Nanostructures and absorption spectra of samples. (a) Schematic of SQD-AgNR nanocomplex. (b) SEM image of the AAO template loaded with AgNRs at the barrier layer side. (c) TEM image of the AgNRs with λLSPR = 700 nm. (d) Absorption spectra (θin = 80°) of AgNR arrays with λLSPR = 650, 670, 700, 710, 755 and 795 nm.
Fig. 2
Fig. 2 TPL spectra of SQD-AgNR nanocomplex with λLSPR = 670 nm and Pexc = 5, 15, 20, 25, 30, 35, 50 and 65 mW. The emission peak shifts from ~665 nm to ~698 nm as the excitation power increases from 5 mW to 65 mW. The edge ~740 nm in TPL spectra is caused by the band-pass filter.
Fig. 3
Fig. 3 TPL spectra of SQD-AgNR nanocomplex with λLSPR = 555, 577, 610, 650, 670, 700, 710, 755, and 795 nm, the excitation power is fixed at 65 mW.
Fig. 4
Fig. 4 Peak emission wavelength of the SQD-AgNR nanocomplex as a function of excitation power. Here, LSPR wavelengths of the AgNR arrays are 577, 610, 650, 670, 700 and 795 nm.
Fig. 5
Fig. 5 Spectral shifting rate RShift of the SQD-AgNR nanocomplex as a function of LSPR wavelength of the AgNRs array.

Equations (3)

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

ΔλΔ λ 0 + R Shift P exc
Δ E Stark P exc ( ω Laser ω LSPR ) 2 + γ 2
R Shift = Δ E Stark P exc = A ( ω Laser ω LSPR ) 2 + γ 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.