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

Random lasing assisted by CuSO4 and Au nanoparticles in random gain systems

Open Access Open Access

Abstract

Random laser actions from a dye pyrromethene-597(PM597)-doped CuSO4 mixture infiltrated in glass capillaries are demonstrated. Discrete lasing modes are observed at the complete solid angle. Low threshold and variable emission wavelength of the random lasers are realized by adjusting the concentration of CuSO4 in the mixtures. With the exception of the multiple scattering of light caused by micro-crystals dispersed in the mixtures, another underlying physical mechanism for tuning random laser is closely related to energy transfer between the molecules of PM597 and CuSO4. Meanwhile, we also demonstrate the realization of a plasmonic random laser by doping Au nanoparticles (NPS) in the PM597-doped CuSO4 mixture. We find that the localized surface plasmon resonance of Au NPS lowers the threshold, increases the emission intensity and reduces the number of the lasing modes of the random lasers. Our results indicate that CuSO4 paves a new avenue towards designing low-cost photonic devices based on the random lasers.

© 2017 Optical Society of America

1. Introduction

Random laser has received considerable research attention since it was proposed by Letokhov in the late 1960s [1]. Unlike traditional laser, the generation of random laser is free of a resonance cavity with high-quality reflection mirrors. The latent cavities for random laser are formed by the feedback of photons due to multiple scattering in disordered gain-scattering media [2–5]. When the scattered photons are amplified efficiently in the latent cavities and make the gain larger than the loss in the system, lasing will occur. The combination between scattering and amplification triggers the random laser. Random laser possesses distinct feature owing to its unique excitation mechanism involving low spatial coherence and simplicity. Hence, random laser has many potential applications, including nanotechnology, full field imaging and projection [6]. The amplified-scattering mechanism has been realized in numerous materials, such as semiconductor powder and those of dye-doped liquid crystals, biomaterials or polymer [4,7–10].

Metal nanoparticles (NPS), due to their unique property of localized surface plasmon resonance (LSPR) and enhanced scattering characteristic, have been widely doped in the random gain media to adjust coherent and incoherent random lasing [11–19]. On the one hand, the LSPR will enhance the localized electromagnetic field around the metal NPS. On the other hand, compared to dielectric scatters, the metal NPS have larger scattering cross section for visible light and can result in enhanced scattering strength within these random media. An incoherent random emission in dye-doped Ag NPS random gain system was firstly reported by Dice et al. [12]. Following the route, Popov et al. showed the enhanced lasing characteristic by embedding Au NPS in a thin polymer film containing dyes [13]. By introducing metal NPS with different material, shape, dimension and aggregation, the effect of metal NPS on random lasing was concretely investigated [14–16]. Recently, Ye et al. studied the variation of threshold, turn-on time and turn-off time of the random laser by varying the concentration of Ag NPS in dye-doped nematic liquid crystals [8]. An ultra-thin plasmonic random laser which consists of a free-standing polymer membrane embedded with Ag NPS was realized by Zhai T.R et al. [17], and they also fabricated a mechanically-tunable random laser based on a waveguide-plasmonic scheme [18]. Hu Z.J et al. reported the random fiber laser based on the LSPR of Au NPS in liquid core optical fiber [19].

Recently, we reported random lasing in a disordered system that made up of manganese chloride and dye-doped polymer dispersed liquid crystals. Low-threshold and high-intensity random laser observed in the system is based on the energy transfer between dye molecules and manganese chloride molecules [20]. As we know, due to the random walk of the photons, many different latent cavities are formed in the random gain systems. So the emission of the random systems is not in a specific direction, it can be observed in complete solid angle. However, reports about this perfect emission are few. In this work, we investigated the random lasing from dye Pyrromethene-597 (PM597) doped supersaturated copper sulfate (CuSO4) capillaries samples through varying the concentration of CuSO4 or adding Au NPS as extra light scatters. The results show that random laser can be observed in all detected direction around the PM597-doped CuSO4 sample, meanwhile, the threshold and energy of random laser can be tuned neatly by varying the CuSO4 concentration. When Au NPS is doped in the sample, the threshold of random laser is lowered and the number of lasing modes is reduced. Our demonstrated approach not only indicates CuSO4 is a favorable scatters in random gain systems but also pave a new avenue towards designing low-cost optical devices based on random laser.

2. Experiments

Pyrromethene-597 (PM597) dispersed in ethanol was chosen as optical gain material in this work. Solid CuSO45H2O (7.354 g) was dispersed into deionized water (10 mL) to form CuSO4 supersaturated solution. We prepared four uniform mixtures of PM597-doped CuSO4, which contained 0.3 mg PM597, 0.2 mL ethanol and different volumes of CuSO4 supersaturated solution (0.1, 0.2, 0.25 and 0.3 mL). For comparison, we added 0.1 mL ethanol solution of Au NPS into the above PM597-doped CuSO4 (with 0.2 mL CuSO4), where the number density of the Au NPS is 1.025×1011/mL. Figure 1(b) illustrates the morphology of spherical Au NPS we used, revealing the average diameter is 50 nm. All the mixtures were shaken for 2 hours by ultrasonic and poured into glass capillaries, as illustrated in Fig. 1(c), which is of 50 mm in length and 0.5 mm in diameter. The ingredient of all the samples is shown in Table 1.

 figure: Fig. 1

Fig. 1 (a) Experiment setup for measuring random lasing. (b) TEM images of Au NPS with average diameter 50 nm. (c) Schematic diagram of the ingredients in sample. (d) The normalized absorption intensity of PM597, Au NPs, CuSO4 and fluorescence intensity of PM597.

Download Full Size | PDF

Tables Icon

Table 1. The ingredient of all the samples.

The experimental scheme for emission measurement is shown in Fig. 1(a). The pump beam from a second-harmonic of a Q-switched Nd: YAG laser (532 nm wavelength, 10 HZ repetition rate and 8 ns pulse duration) was divided into two sub-beams by a polarizing beam splitter. The reflected sub-beam emission is collected as a reference beam by energy meter while the transmitted sub-beam, after propagating a half-wave plate, is focused on the capillary samples by a lens and a slit. Emission from the capillary sample’s surface was collected with an optical fiber bundled into a spectrometer to resolve the spectra.

3. Results and discussion

Figure 2(a) plots the evolution of emission spectra with pump energy for PM597 system. Broad fluorescence spectra with a full width at half maximum (FWHM) of about 20 nm can be observed under different pump energies. The fluorescence intensity is enhanced with the increment of the pump energy. The evolution of emission spectra against pump energy for PM597-doped CuSO4 (0.1 mL) is shown in Fig. 2(b). At low pump energy, there is only a broad spontaneous emission spectrum centered at 604.8 nm with the FWHM of about 7.5 nm. A single sharp peak centered at 607.7 nm with the FWHM of about 0.23 nm over the broad background spectrum appears when the pump energy increases to 9.4μJ. As the pump energy increases further to 16.4μJ, multiple sharp peaks emerge with the FWHM much less than 0.20 nm. Moreover, the number of the sharp peaks increases with the pump energy increasing, indicating the occurrence of coherent random lasing. This is closely related to CuSO4 micro-crystal molecules that not fully dispersed in the mixture sample, which causes multiple scattering of photons in the sample. Due to the multiple scattering and wave interference, the photons may be trapped inside the gain medium and form different optical feedback loops of photons. The photons are amplified in the optical feedback loop and exceed the loss, implying the occurrence of a random laser. Because of the random distribution of scattering particles inside the PM597-CuSO4 mixture, the optical feedback loops of photons in it will be random, resulting in multiple lasing modes. When the pump energy increases, gain greater than loss appears in more optical feedback loop, which results in more discrete peaks in the emission spectrum. In addition, as shown in Fig. 1(d), the emission spectrum of PM597 overlaps with the absorption spectrum of CuSO4, which indicates that the random lasing phenomenon in the PM597 doped CuSO4 (0.1 mL) is also associated with the energy transfer between CuSO4 molecules and PM597 molecules. The photons emitted by dye PM597 is subsequently absorbed by CuSO4 molecules. Besides, the dye molecules may directly transfer its excitation energy to a ground-state of the CuSO4 molecules [21]. The pump-dependence of the integrated emission intensities for PM597, as depicted in Fig. 2(c), do not exhibit threshold behavior, implying spontaneous emission of PM597 molecules. However, the pump-dependence of the integrated emission intensity in PM597-doped CuSO4 (0.1 mL) shows a clear threshold behavior, which is a typical characteristics involved in random lasing with coherent feedback. Figure 2(d) plots single-shot emission spectra of PM597 doped CuSO4 (0.1 mL) system at the pump energy of 16.65μJ, where the laser spikes as well as emission intensities exhibit chaotic behavior, indicating pump-dependent competitions among different laser multimodes.

 figure: Fig. 2

Fig. 2 The evolution of emission spectra of (a) PM597 and (b) PM597-doped CuSO4 (0.1 mL) at different pump energies. (c) Dependence of emission intensity on the pump energies for PM597 and PM597-doped CuSO4 (0.1 mL). (d) Single shot lasing spectra of PM597-doped CuSO4 (0.1 mL) at the pump energy of 16.65 μJ.

Download Full Size | PDF

Figure 3(a) depicts emission spectra versus different detection angles θ (see Fig. 1(c)) for PM597-doped CuSO4 (0.25 mL) system. It is found that lasing modes are observed at the complete solid angle, however, lasing modes are different at different angles. Since the different optical feedback loops could have different output directions, and may have stronger emission into some directions. The peak intensity of lasing signal as a function of detection angle is presented in Fig. 3(b), it is noted that the intensity of lasing signal is highly dependent on the detection angle and reaches the maximum at the angle of θ=900. This is reasonable since the angle of θ=900 is the incident direction of the pump light, corresponding to the most efficient illumination of pump light. Moreover, in this direction light may experience the longest optical path of gain and can be absorbed much more fully than in other directions, reaching the strongest output intensity of random lasing.

 figure: Fig. 3

Fig. 3 (a) Emission spectra of PM597-doped CuSO4 (0.25 mL) recorded at different detection angles. (b) Dependence of peak emission intensity on the detection angles for PM597-doped CuSO4 (0.25 mL).

Download Full Size | PDF

Single-shot emission spectra of PM597-doped CuSO4 with the concentration of CuSO4 0.1 mL, 0.2 mL and 0.3 mL are illustrated in the inset of Fig. 4(b), where the samples are pumped with the same pump energy of 12.00μJ. Red shift of the main peak wavelength is observed with increasing the concentration of CuSO4 from 0.1 mL to 0.3 mL. It is responsible for the thicker density of CuSO4 that supports, for relatively redder photons, more efficient energy transfer between the dye PM597 molecules and those of CuSO4. In particular, we find that the spectrum of PM597 doped CuSO4 (0.3 mL) appears two main peaks centered at 611.79 and 627.28 nm respectively, which seems to be a manifestation of energy level splitting in the coupled-resonance transition of dye and CuSO4 molecules.

 figure: Fig. 4

Fig. 4 (a) Emission intensity as a function of the pump energy for PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL). (b) Ensemble-averaged Power Fourier Transform (PFT) analysis for single-shot emission spectra (sum = 50) of PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL) at the pump energy of 12.00μJ. Inset: Two individual single-shot emission spectra recorded for the PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL) at the pump energy of 12.00μJ.

Download Full Size | PDF

Figure 4(a) plots the dependence of integrated emission intensities upon the pump energies for the three PM597-doped CuSO4 mixture samples. All the samples manifest threshold behavior and have thresholds of 8.9, 7.2 and 9.25μJ, corresponding to concentrations of CuSO4 0.1, 0.2 and 0.3 mL, respectively. The sample with CuSO4 0.2 mL has the lowest threshold, which is attributed to the highest cavity quality factor of the random laser in it. The quality factor Q is a parameter defined commonly to exhibit the peculiarity of laser cavity, the higher value it is, the less loss the resonant cavity will be. For mirrorless resonant cavity of random laser, the quality factor is given by [22,23]

Q=2cnL3εm+Re(εsεm)l(ckDlnR)εmRe(εsεm).
Here, where n is the refraction index of medium, L and l correspond to the resonant cavity length in random laser and the characteristic size of active particles, c stands for the light speed in vacuum, εs and εm represent the permittivity of scatters and host medium, D is the extinction coefficient and R is the reflection of resonant cavity. Therefore, the highest cavity quality factor means the longest resonant cavity length Lof random laser.

To further verify the above judgement, we calculated the ensemble-averaged Power Fourier Transform (PFT) analysis (see Fig. 4(b)) for a sum of single-shot emission spectra of the three PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL) mixture samples, respectively. All the single-shot emission spectra are collected at various sample positions upon the same pump energy of 12.00 μJ. Positions of the peaks in PFT spectra are expressed as follows [24,25],

dm=nLm/π.
where m and L are an integer denoting the Fourier transform harmonics and oscillation cavity length of a random laser, respectively. As a consequence, based on the first peak position in the PFT spectra (m=1 and n=1.46), we derived the cavity lengths of the random lasers to be, 3.81, 8.52 and 1.72 μm, corresponding to the samples with CuSO4 0.1, 0.2 and 0.3 mL, respectively. Thus, the longest cavity length exactly supports the lowest threshold of the random laser. As we know, the doping density of CuSO4 changes the occurrence frequency of scattering events and it is harder for photons to escape through the surface of the scatters for the sample with CuSO4 0.2 mL. Consequently, the sample with CuSO4 0.2 mL belongs to the optimum configuration of scattering particles, which means the highest quality cavities [26].

Figure 5 demonstrates the effect of the pump energy on random lasing spectra and pump–emission intensity dependence for the samples PM597-doped CuSO4 (0.2 mL) without and with Au NPS. For PM597-doped CuSO4 (0.2 mL), the emission spectra of random laser at different pump energy is shown in Fig. 5(a), where a random lasing main peak centered at 610.2 nm with the FWHM of about 0.21 nm emerges at the pump energy of 7.65μJ. As a comparison, random lasing action of PM597-doped CuSO4 (0.2 mL) with Au NPS is depicted in Fig. 5(b), which exhibits a sharp peak located at 602.1 nm with the FWHM of about 0.23 nm at the pump energy of 5.96μJ. Figure 5(c) plots the evolution of the integrated emission intensities of the output random lasers with pump energies, corresponding to the PM597-doped CuSO4 (0.2 mL) without and with Au NPS, respectively. It shows in the both systems that the emission intensities increase gradually with the increase of the pump energies, where an abrupt increase of the emission intensities occurs, suggesting the threshold behavior of the both systems. In addition, thresholds of the former and the latter systems are 7.20 and 5.26μJ, respectively, indicating that the threshold is reduced as doping Au NPS in the PM597-doped CuSO4 (0.2 mL) sample. This is owing to the localized surface plasmon resonance (LSPR) and the larger scattering cross section caused by Au NPS. The LSPR enhances the localized electromagnetic field around the Au NPS which facilitates the simulated emission of laser dye, while the larger scattering cross section provides stronger multiple scattering of light which result in longer resonance cavity length of random laser and reduces the pumping threshold.

 figure: Fig. 5

Fig. 5 The evolution of emission spectra of PM597-doped CuSO4 (0.2 mL) (a) without and (b) with Au NPS at different pump energies. (c) Dependence of emission intensity on the pump energies for PM597-doped CuSO4 (0.2 mL) without and with Au NPS. (d) Ensemble-averaged PFT analysis for single-shot emission spectra (sum = 50) of PM597-doped CuSO4 (0.2 mL) without and with Au NPS at the pump energy of 12.75μJ. Inset: Two individual single-shot emission spectra recorded for the PM597- doped CuSO4 (0.2 mL) without and with Au NPS at the pump energy of 12.75μJ.

Download Full Size | PDF

Similar to Fig. 4(b), ensemble-averaged PFT curves for a sum of (sum = 50) single-shot emission spectra of PM597-doped CuSO4 (0.2 mL) without and with Au NPS at the same pump energy of 12.75μJare plotted in Fig. 5(d). The inset of Fig. 5(d) depicts two individual single-shot emission spectra of PM597-doped CuSO4 (0.2 mL) without and with Au NPS at the same pump energy of 12.75μJ. The single-shot emission spectral of the sample with Au NPS appears a clear blue-shift with respect to that of the sample without Au NPS, which is due to the variance of fluorescence quenching efficiency caused by the LSPR of Au NPS. In addition, it is worth noting that the number of lasing modes in the sample with Au NPS is less than that in the sample without Au NPS. This can also be observed in the inset of Fig. 5(d), in which the PFT curve of the sample with Au NPS has fewer number of the Fourier transform harmonics. When Au NPS is doped in the sample, a longer scattering mean free path results in less number of resonant cavities where the lasing threshold can be reached [27]. According to Fig. 5(d), we obtain the oscillation cavity lengths of the samples without (8.60μm) and with Au NPS (26.23μm). Evidently, the resonant cavity length of the sample is increased by three times when Au NPS is doped in it, the longer cavity length means higher quality factor of the cavity length or lower threshold of random laser. Besides, the partial overlap between absorption spectrum of Au NPS (d = 50 nm) and fluorescence emission spectrum of PM597 dyes, as plotted in Fig. 1(d), implies that the photon emitted by PM597 dye molecule can be absorbed by Au NPS. This mechanism can also enhance the LSPR effect of Au NPS and extends the lifetime of laser photons in the system.

Contour maps of shot-to-shot lasing spectra at different pump pulses for PM597-doped CuSO4 (0.2 mL) without and with Au NPS are depicted in Fig. 6(a) and 6(b) while keeping pump position and pump energy (13.10μJ) fixed. It is seen that the lasing in the former system (Fig. 6(a)) shows unstable frequencies but the lasing frequencies in the latter system (Fig. 6(b)) are more stable, moreover, the latter system shows higher intensity. The variable lasing frequencies in the former system is correlated to the dynamic competition of multiple laser modes. However, for the systems with Au NPS, the enhanced LSPR can increase the coherence of optical field, inducing steadier lasing modes and fewer laser spikes (see Fig. 5(d)).

 figure: Fig. 6

Fig. 6 Contour maps of shot-to-shot lasing spectra at different pump pulses for PM597-doped CuSO4 (0.2 mL) (a) without and (b) with Au NPS at the pump energy of 13.10μJ.

Download Full Size | PDF

4. Conclusion

To summarize, we studied the random lasing from PM597-doped CuSO4 glass capillaries. Results show that the threshold, lasing modes, peak intensity and emission wavelength can be tuned by changing the concentration of CuSO4 in the mixture materials. Diverse emission spectra are observed in different detection directions around the surface of the sample. This is responsible for the multiple scattering of CuSO4 micro-crystal dispersed in the mixture materials. In addition, the energy transfer between dyes molecules and CuSO4 also plays an important role in the tuning properties of the random laser. The comparison of cavity lengths in the three systems confirms the effect of CuSO4 on the random lasing. Besides, random laser with lower threshold, higher intensity and fewer modes was found when Au NPS is doped into the PM597-doped CuSO4 sample, which is due to LSPR and a larger scattering cross section provided by Au NPS. The PFT-analysis on laser cavity lengths and the contour maps on time-frequency domain of shot-to-shot lasing spectra for both systems reveal the positive influence of Au NPS on improving random laser performance. These results not only enrich the realization schemes of general random laser but also provide insights into rules of random lasing occurrence in some complex gain media with energy transfer between their ingredients.

Funding

National Natural Science Foundation of China (11474021).

References and links

1. V. S. Letokhov, “Generation of light by a scattering medium with negative resonance absorption,” Sov. Phys. JETP 26(4), 835 (1968).

2. N. M. Lawandy, R. M. Balachandran, A. S. L. Gomes, and E. Sauvain, “Laser action in srongly media,” Nature 368(6470), 436–438 (1994). [CrossRef]  

3. J. Andreasen, A. A. Asatryan, L. C. Botten, M. A. Byrne, H. Cao, L. Ge, L. Labonté, P. Sebbah, A. D. Stone, H. E. Türeci, and C. Vanneste, “Modes of random lasers,” Adv. Opt. Photonics 3(1), 88–127 (2011). [CrossRef]  

4. H. Cao, Y. G. Zhao, S. T. Ho, E. W. Seelig, Q. H. Wang, and R. P. H. Chang, “Random laser action in semiconductor powder,” Phys. Rev. Lett. 82(11), 2278–2281 (1999). [CrossRef]  

5. C. R. Lee, S. H. Lin, J. W. Guo, J. D. Lin, H. L. Lin, Y. C. Zheng, C. L. Ma, C. T. Horng, H. Y. Sun, and S. Y. Huang, “Electrically and thermally controllable nanoparticle random laser in a well-aligned dyedoped liquid crystal cell,” Opt. Mater. Express 5(6), 1469–1481 (2015). [CrossRef]  

6. D. S. Wiersma, “The physics and applications of random lasers,” Nat. Phys. 4(5), 359–367 (2008). [CrossRef]  

7. K. L. van der Molen, R. W. Tjerkstra, A. P. Mosk, and A. Lagendijk, “Spatial extent of random laser modes,” Phys. Rev. Lett. 98(14), 143901 (2007). [CrossRef]   [PubMed]  

8. L. H. Ye, B. Liu, F. J. Li, Y. Y. Feng, Y. P. Cui, and Y. Q. Lu, “The influence of Ag nanoparticles on random laser from dye-doped nematic liquid crystals,” Laser Phys. Lett. 13(10), 105001 (2016). [CrossRef]  

9. R. C. Polson and Z. V. Vardeny, “Random lasing in human tissues,” Appl. Phys. Lett. 85(7), 1289–1291 (2004). [CrossRef]  

10. D. F. Huang, M. Xu, X. Y. Liu, M. Yang, T. Yi, C. K. Wang, T. S. Li, and S. Y. Liu, “Low threshold random lasing actions in natural biological membranes,” Laser Phys. Lett. 13(6), 065603 (2016). [CrossRef]  

11. S. Y. Ning, H. Dong, N. M. Zhang, J. N. Zhao, and L. Ding, “Plasmonic enhancement of random lasing from dye-doped polymer film by bristled Ag/TiO2 composite nanowires,” Opt. Mater. Express 6(12), 3725–3732 (2016). [CrossRef]  

12. G. D. Dice, S. Mujumdar, and A. Y. Elezzab, “Plasmonically enhanced diffusive and subdiffusive metal nanoparticles-dye random laser,” Appl. Phys. Lett. 86(13), 131105 (2005). [CrossRef]  

13. O. Popov, A. Zilbershtein, and D. Davidov, “Random lasing form dye-gold nanoparticles in polymer films: enhanced at the surface-plasmon-resonance wavelength,” Appl. Phys. Lett. 89(19), 191116 (2006). [CrossRef]  

14. X. Meng, K. Fujita, Y. Zong, S. Murai, and K. Tanaka, “Random lasers with coherent feedback from highly transparent polymer films embedded with silver nanoparticles,” Appl. Phys. Lett. 92(20), 201112 (2008). [CrossRef]  

15. X. Meng, K. Fujita, S. Murai, and K. Tanaka, “Coherent random lasers in weakly scattering polymer films containing silver nanoparticles,” Phys. Chem. A 79(5), 053817 (2009).

16. X. Meng, K. Fujita, S. Murai, T. Matoba, and K. Tanaka, “Plasmonically controlled lasing resonance with metallic-dielectric core-shell nanoparticles,” Nano Lett. 11(3), 1374–1378 (2011). [CrossRef]   [PubMed]  

17. T. Zhai, Z. Xu, X. Wu, Y. Wang, F. Liu, and X. Zhang, “Ultra-thin plasmonic random lasers,” Opt. Express 24(1), 437–442 (2016). [CrossRef]   [PubMed]  

18. T. R. Zhai, J. Chen, L. Chen, J. Y. Wang, L. Wang, D. H. Liu, S.T. Li, H. M. Liu, and X. P. Zhang, “A plasmonic random laser tunable through stretching silver nanowaires embedded in a flexible substrate,” Nanoscale 7(6), 2235–2240 (2015).

19. Z. J. Hu, Y. Y. Liang, K. Xie, P. F. Gao, D. G. Zhang, H. M. Jiang, F. Shi, L. C. Yin, J. G. Gao, H. Ming, and Q. J. Zhang, “Gold nanoparticles-based plasmonic random fiber laser,” J. Opt. 17(3), 035001 (2015). [CrossRef]  

20. Z. Z. Shang, M. C. Yang, and L. G. Deng, “Low-Threshold and High Intensity Random Lasing Enhanced by MnCl2,” Materials (Basel) 9(9), 725 (2016). [CrossRef]  

21. L. Cerdán, E. Enciso, V. Martin, J. Banuelos, I. Lopez-Arbeloa, A. Costela, and I. Garcia-Moreno, “FRET-assisted laser emission in colloidal suspensions of dye-doped latex nanoparticles,” Nat. Photonics 6(9), 623–626 (2012). [CrossRef]  

22. V. V. Prosentsov, “Light Amplification and Scattering by Clusters Made of Small Active Particles: The Local Perturbation Approach,” Physical Science International Journal 8(4), 1–11 (2015). [CrossRef]  

23. V. Barna, R. Caputo, A. De Luca, N. Scaramuzza, G. Strangi, C. Versace, C. Umeton, R. Bartolino, and G. N. Price, “Distributed feedback micro-laser array: helixed liquid crystals embedded in holographically sculptured polymeric microcavities,” Opt. Express 14(7), 2695–2705 (2006). [CrossRef]   [PubMed]  

24. D. Hofstetter, L. T. Romano, R. L. Thornton, D. P. Bour, and N. M. Johnson, “Characterization of intra-cavity reflections by Fourier transforming spectral data of optically pumped InGaN lasers,” Appl. Phys. Lett. 71(22), 3200–3202 (1997). [CrossRef]  

25. R. C. Polson, G. Levina, and Z. V. Vardeny, “Spectral analysis of polymer microring lasers,” Appl. Phys. Lett. 76(26), 3858–3860 (2000). [CrossRef]  

26. A. L. Burin, M. A. Ratner, H. Cao, and R. P. H. Chang, “Model for a random laser,” Phys. Rev. Lett. 87(21), 215503 (2001). [CrossRef]   [PubMed]  

27. H. Cao, J. Y. Xu, S. Chang, and S. T. Ho, “Transition from amplified spontaneous emission to laser action in strongly scattering media,” Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics 61(2), 1985–1989 (2000). [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 (6)

Fig. 1
Fig. 1 (a) Experiment setup for measuring random lasing. (b) TEM images of Au NPS with average diameter 50 nm. (c) Schematic diagram of the ingredients in sample. (d) The normalized absorption intensity of PM597, Au NPs, CuSO4 and fluorescence intensity of PM597.
Fig. 2
Fig. 2 The evolution of emission spectra of (a) PM597 and (b) PM597-doped CuSO4 (0.1 mL) at different pump energies. (c) Dependence of emission intensity on the pump energies for PM597 and PM597-doped CuSO4 (0.1 mL). (d) Single shot lasing spectra of PM597-doped CuSO4 (0.1 mL) at the pump energy of 16.65 μJ .
Fig. 3
Fig. 3 (a) Emission spectra of PM597-doped CuSO4 (0.25 mL) recorded at different detection angles. (b) Dependence of peak emission intensity on the detection angles for PM597-doped CuSO4 (0.25 mL).
Fig. 4
Fig. 4 (a) Emission intensity as a function of the pump energy for PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL). (b) Ensemble-averaged Power Fourier Transform (PFT) analysis for single-shot emission spectra (sum = 50) of PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL) at the pump energy of 12.00 μJ . Inset: Two individual single-shot emission spectra recorded for the PM597-doped CuSO4 (0.1, 0.2 and 0.3 mL) at the pump energy of 12.00 μJ .
Fig. 5
Fig. 5 The evolution of emission spectra of PM597-doped CuSO4 (0.2 mL) (a) without and (b) with Au NPS at different pump energies. (c) Dependence of emission intensity on the pump energies for PM597-doped CuSO4 (0.2 mL) without and with Au NPS. (d) Ensemble-averaged PFT analysis for single-shot emission spectra (sum = 50) of PM597-doped CuSO4 (0.2 mL) without and with Au NPS at the pump energy of 12.75 μJ . Inset: Two individual single-shot emission spectra recorded for the PM597- doped CuSO4 (0.2 mL) without and with Au NPS at the pump energy of 12.75 μJ .
Fig. 6
Fig. 6 Contour maps of shot-to-shot lasing spectra at different pump pulses for PM597-doped CuSO4 (0.2 mL) (a) without and (b) with Au NPS at the pump energy of 13.10 μJ .

Tables (1)

Tables Icon

Table 1 The ingredient of all the samples.

Equations (2)

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

Q= 2cnL 3 ε m +Re( ε s ε m ) l( ckDlnR ) ε m Re( ε s ε m ) .
d m = nLm /π .
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.