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

An optimization procedure for the design of all-optical switches based on metal-dielectric nanocomposites

Open Access Open Access

Abstract

We present a procedure to optimize the performance of all-optical switches based on metal-dielectric nanocomposites. The management of constructive and destructive interference between the third-, fifth- and seventh-order susceptibilities allowed characterization of optimal conditions for ultrafast switching with reduced losses. Proof-of-principle experiments with metal-colloids are reported to validate the method.

© 2015 Optical Society of America

1. Introduction

Ultrafast all-optical switches (AOS) are essential components of high-speed photonic networks and the efficient control of light-by-light by using nonlinear (NL) media is of great interest nowadays [1–5]. For selection of materials with proper third-order susceptibility, χ(3), for AOS, two figures-of-merit are considered: T=|α2λ(n2)1| and W=|Δn(α0λ)1|, where λ is the laser wavelength, α0 is the linear absorption coefficient, α2Imχ(3) is the two-photon absorption coefficient, n2Reχ(3) is the third-order refractive index and Δn is the maximum variation of the refractive index that can be induced in the material. T<1 and W>1 are required to obtain efficient AOS; hence, NL materials should present fast response, high transmittance, large NL refractive index and small NL absorption coefficient. Unfortunately, large n2 normally corresponds to large α2 [6], and therefore it becomes difficult to find proper materials for efficient AOS [7]. Also, one should pay attention that the highly NL materials available present high-order nonlinearities (HON) that may affect the evaluation based on the third-order susceptibility only.

Among the physical systems considered for AOS the metal-dielectric nanocomposites (MDNCs) with metallic nanoparticles (NPs) deserve special attention due to their high optical susceptibility, ultrafast response and the possibility of changing their NL susceptibility by changing the NPs volume fraction, f – the ratio between the volume occupied by the NPs and the host [8–12]. For instance, it was reported that gold NPs embedded inside an oxide glass film improved T by more than one-order of magnitude at 800 nm [10]. Limitations for use of MDNCs are the large linear optical absorption at the localized surface plasmons (LSP) resonance frequency and strong NL absorption [13–16]. In principle the optical response of MDNCs varies with f [14], the NPs environment [17, 18], the size and shape of NPs [19, 20], and the laser wavelength [20]. The NL response of MDNCs is described by effective susceptibilities that contain information on the NPs and the host material. Enhancement, decrease and even suppression of specific NL susceptibilities of MDNCs can be obtained varying f and the incident laser intensity [21–23].

In this work we applied the nonlinearity management procedure of [21–23] in order to demonstrate the optimization of AOS in proof-of-principle experiments with metal-colloids. Section 2 describes the samples preparation with silver (Ag) NPs suspended in acetone and experimental details. Section 3 presents the NL characterization of samples with various NPs volume fractions for different laser intensities. Considering the measured NL parameters it is shown that the nonlinearity of MDNCs can be controlled to obtain figures-of-merit for all-optical switching that may be enhanced by about two-orders of magnitude. Section 4 presents a summary of the results.

2. Experimental details

Initially colloids containing Ag NPs suspended in water were prepared following the method presented in [24]. 90 mg of AgNO3 were diluted in 500 ml of water at 100°C; 10 ml of 1% sodium citrate solution was added for reduction the Ag+ ions, and later was boiled and strongly stirred for 1 h. A colloid with Ag NPs of various shapes was obtained. Subsequently, photofragmentation of the NPs was performed by irradiation of the pristine colloid, under slow stirring, using the second harmonic beam at 532 nm obtained from a Nd: YAG laser (8 ns, 85 mJ/pulse, 10 Hz) for 1 h, according to [25]. The photofragmentation of the NPs is due to their melting and vaporization because of the large absorption of the laser energy by the particles and low heat-transfer for the hosting medium [26, 27]. After photofragmentation metal-colloids with 0.5×105f2.5×104 were obtained by adding 20 µl to 300 µl of the Ag-water suspension in 1 ml of acetone. The Ag NPs do not aggregate due to the sodium citrate molecules attached to their surface; the shape and size of the NPs remained unchanged for at least 3 months.

The linear absorption spectra were measured using a commercial spectrophotometer. The light source for the NL measurements was the second harmonic beam at 532 nm of a Q-switched and mode-locked Nd: YAG laser that delivers selected single pulses of 80 ps with repetition rate of 10 Hz.

Z-scan [28] and Kerr shutter [29] experiments were performed to characterize the samples. In the Z-scan setup the laser beam was focused by a 10 cm focal length lens (beam waist: 20 µm) on a sample with thickness of 1 mm, contained in a quartz cell. Photodetectors placed in the far-field region, with adjustable apertures in front of them, were used to measure the intensity transmitted by the sample in different Z positions along the beam propagation direction. Closed-aperture (CA) and open-aperture (OA) Z-scan schemes were used. A reference channel was used to improve the signal-to-noise ratio as in [30]. For the Kerr shutter setup the laser beam was split into probe and pump beams with relative intensities: Iprobe=0.1(Ipump). The angle between the beams was 2.3° and the angle between their electric fields was 45°. Both beams were focused inside the sample by a 10 cm focal length lens. When the two beams overlap spatially and temporally inside the sample, the pump beam induces a NL birefringence that induces rotation of the probe beam electric field. Then, a fraction of the transmitted Iprobe by the sample passes through a polarizer crossed to the input probe beam electric field. A detector was used to record the transmitted Iprobe versus the delay time, τ, between the pump and probe pulses. Liquid carbon disulfide (CS2) with n2=+3.1×1014cm2/W [28] was the reference standard for calibration of the measurements.

All NL experiments were repeated more than one time with each sample and the results were reproduced.

3. Results and discussion

Figure 1(a) shows the linear absorbance spectra of the metal-colloids before (dotted line) and after (solid line) photofragmentation of the original NPs synthesized. The smaller LSP resonance linewidth exhibited by the laser-ablated colloid indicates a homogeneous distribution of Ag NPs sizes. The absorbance spectrum of pure acetone (dashed line) presents large transparency window, corroborating that the resonance at ~400 nm is due to the LSP in the Ag NPs. Figure 1(b) shows the size distribution histogram of the spherical NPs with average diameter of 9 nm, obtained using a Transmission Electron Microscope (TEM).

 figure: Fig. 1

Fig. 1 (a) Linear absorption spectra of the metal-colloid with f=4.0×105 and acetone (cell thickness: 1 mm). (b) Size distribution histogram of the NPs after photofragmentation. A TEM image of the silver NPs is shown in the inset.

Download Full Size | PDF

Figure 2 shows the normalized Z-scan profiles for f=2.0×105, 3.7×105, 5.9×105 and 8.0×105, with the intensity in the focus adjusted to be I0=9.5GW/cm2. The Z-scan profiles in Fig. 2(a) display a defocusing nonlinearity with small features on the third-order profiles due to HON [21, 22]. Figure 2(b) shows profiles that indicate saturated absorption due to the small detuning between the laser and the LSP frequencies.

 figure: Fig. 2

Fig. 2 Normalized (a) closed-aperture and (b) open-aperture Z-scan profiles obtained for different NPs volume fractions. Laser peak intensity: 9.5GW/cm2.

Download Full Size | PDF

The NL response of the metal-colloid is described by effective susceptibilities, χeff(2N+1), N=1,2,3,, that contain information on the NL susceptibilities of the host liquid and the NPs. The total NL susceptibility with contributions up to the seventh-order is given by χNL(|E|)=34χeff(3)|E|2+58χeff(5)|E|4+3564χeff(7)|E|6 [31], where E is the amplitude of the laser electric field. χeff(2N+1), N=13, were calculated in [22] assuming that the NPs diameters are smaller than their relative distances that are smaller than λ. The NL measurements allow to determine n2NReχeff(2N+1) and α2NImχeff(2N+1) as in [21, 22].

The NL refractive indices were obtained from the CA Z-scan data, plotting the normalized peak-to-valley transmittance change per incident intensity, |ΔTPV|/I, versus I and making a polynomial fit using the expression

ΔTPVI0.396kLeff(1)n2+0.198kLeff(2)n4I+0.102kLeff(3)n6I2,
where k=2πn0/λ, Leff(N)=(1eNα0L)/(Nα0) and L is the sample length.

The NL absorption coefficients were obtained from the OA Z-scan experiments fitting the expression

ΔTI(2)32Leff(1)(α2+α4I+α6I2).

The solid lines in Fig. 2 correspond to the normalized laser transmittance in the far-field as a function of the sample's position given by

T(z,ΔΦ0)1+N=13(4N)ΔΦ0(2N+1)z/z0[(z/z0)2+(2N+1)2][(z/z0)2+1]N,
in the CA Z-scan scheme. The lines corresponding to the OA scheme were obtained using the expression
T(z,q0)1πq0ln[1+q0exp(τ2)]dτ.
The on-axis phase-shift at the beam focus is given by ΔΦ0(2N+1)=kn2NIN[1exp(Nα0L)]/(Nα0); q0=αNLILeff(1)/(1+z2/z02) [28] and z0 is the Rayleigh length of the focused beam.

Figures 3(a) and 3(b) show the linear dependence of the NL refractive indices and the NL absorption coefficients with f for I0=9.5GW/cm2. The linear dependence with f was also measured for I0=5.0, 8.6, and 10GW/cm2but is not presented. Therefore, the total effective NL refractive index nNL(I)=n2+n4I+n6I2 and the total effective NL absorption coefficient αNL(I)=α2+α4I+α6I2 also vary linearly with f as shown in Fig. 3(c) and Fig. 3(d). The solid lines were obtained using the generalized Maxwell-Garnett (GMG) model [22] as discussed below.

 figure: Fig. 3

Fig. 3 Dependence with the NPs volume fraction of: (a) NL refractive indices, (b) NL absorption coefficients, (c) total effective NL refractive index and (d) total effective NL absorption coefficient. In (a) and (b) the laser peak intensity was 9.5GW/cm2. Normalized (e) Closed-aperture and (f) Open-aperture Z-scan profiles for f=5.9×105, obtained for different laser peak intensities. The solid lines were obtained from Eqs. (3) and (4).

Download Full Size | PDF

Figures 3(e) and 3(f) show the Z-scan profiles for the same intensities as in Figs. 3(c) and 3(d) with f=5.9×105. The solid curves were obtained from Eqs. (3) and (4) using the NL coefficients found in Figs. 3(a) and 3(b). Notice that |nNL(I)| increases while increasing the laser intensity; on the other hand |αNL(I)| reaches a minimum value for I0=8.6GW/cm2. Therefore, the results of Fig. (3)-(5) show that using the nonlinearity management procedure of [21–23], which consists in controlling the NL response of a MDNC by selecting appropriate values of f and I, it is possible to obtain large nNL(I) and small αNL(I) simultaneously.

 figure: Fig. 4

Fig. 4 (a) Transmitted Kerr signal in 532 nm for silver colloid for pump intensities of: I1=10GW/cm2, I2=8.6GW/cm2, I3=8.0GW/cm2 and I4=7.0GW/cm2 and Iprobe=0.1(Ipump). The inset is the Kerr signal for CS2. (b) Dependence of |ΔT|/Ipump as a function of the pump intensity. Volume fraction: 5.9×105.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Figures-of-merit for all-optical switching.

Download Full Size | PDF

The origin of the effective NL susceptibilities of the metal-colloid was discussed in [13–18, 32, 33]. For the laser wavelength and pulse duration used the NL response is mainly due to the free electrons in the NPs. The influence of thermal effects is negligible due to the low pulses repetition rate. The dependence of the effective NL parameters with f and with the materials susceptibilities is understood using the GMG model [22] which describes the total NL susceptibility of the colloid in terms of the host third-order susceptibility, χh(3), and the NPs susceptibilities, χnp(2N+1), N=13.

Expressions for the effective third-, fifth- and seventh-order susceptibilities are obtained as

χeff(3)=fL2|L|2χnp(3)+χh(3),
χeff(5)=fL2|L|4χnp(5)610fL3|L|4(χnp(3))2310fL|L|6|χnp(3)|2,
χeff(7)=fL2|L|6χnp(7)+1235fL4|L|6(χnp(3))3+335f|L|8[4L2χnp(3)+|L|2(χnp(3))*]|χnp(3)|247fL|L|6[2L2χnp(3)+|L|2(χnp(3))*]χnp(5),
where L=3εh(L)/(εnp(L)+2εh(L)) and εnp(L) (εh(L)) is the linear dielectric function of the NPs (host).

The value of χeff(3) is affected by the competition between χnp(3) and χh(3), that have opposite signs; as a result χeff(3) changes sign as a function of f [22]. On the other hand, the contributions of L and χnp(2N+1), N=13, for the susceptibilities χeff(5) and χeff(7) are more complex as can be observed from Eqs. (6)-(7).

Figure 4(a) shows the Kerr shutter signals for f=5.9×105 with pump intensities equal to 10, 8.6, 8.0, and 7.0GW/cm2. The symmetric profiles are due to the fast NL response of the free electrons in the NPs. Similar behavior was observed for other values of f. The inset of Fig. 4(a) shows the CS2 signal that exhibits asymmetry with respect to τ due to molecular reorientation. The result for CS2 is shown to illustrate the fast temporal response of the apparatus. The signal for pure acetone is smaller than our detection limit.

Figure 4(b) shows a polynomial dependence of ΔT/Ipump versus Ipump (red line) due to HON. For Ipump4GW/cm2 the ratio ΔT/Ipump is constant and the signal is properly described by n2. For 4GW/cm2<Ipump7GW/cm2 the ratio ΔT/Ipump presents linear dependence with Ipump and the slope of the straight line is related to n4. For Ipump>7GW/cm2 the contribution of n6 becomes relevant as already observed in [22].

Figure 5 summarizes the results for W and T but, for better visualization of the results we plotted T1 instead of T. In the present case n2 should be replaced by nNL(I) and instead of α2 we consider αNL(I). Figure 5(a) shows W versus f for intensities between 7.0 and 10GW/cm2. The dependence of Δn and α0 with f was considered. For example for f=5.9×105 the values of Δn≈10−4 and α0=0.06mm-1 were used. Then, W>1 was obtained for intensities larger than 7.0GW/cm2. Figure 5(b) shows T1 versus f. Notice, for instance, that for 10 GW/cm2 we obtained T13 corresponding to f>2×105, while for 8.6 GW/cm2 the value of T1 increases by about two orders of magnitude.

For a more detailed evaluation, Table 1 gives the values of the NL parameters for particular choices of f, laser intensity and the corresponding figures-of-merit for all-optical switching. Notice that the values of nNL increase by approximately a factor of 3 for an increase of f from 3.0×105 to 8.0×105, for a constant intensity. However, a dramatic variation of αNL, which produces a significant reduction of the figure of merit T, is observed for small variations in intensity and fixed value of f. The very small values of αNL, for I0=8.6GW/cm2, were obtained due to the destructive interference between the contributions of negative α2 and α6 with positive α4. Therefore, metal-colloids with appropriate values of f and laser intensity may present optimal figures-of-merit for all-optical switching.

Tables Icon

Table 1. NPs volume fraction, f, laser intensity, I, and the corresponding total NL refractive indices, nNL(I), total NL absorption coefficients, αNL(I), and figures-of-merit (T and W)

Clearly the results of Fig. 5 and Table 1 indicate that the nonlinearity management procedure applied here can be much useful for the fabrication of efficient AOS.

4. Conclusions

In summary, we reported a procedure to fabricate appropriate MDNCs for AOS by managing the effective NL susceptibility. Fast NL response and optimized values of the figures-of-merit for all-optical switching were obtained by adjusting the NPs volume fraction and the incident laser intensity. The present results will be useful for the design of AOS based on MDNCs if proper manipulation of host material and metal NPs is made.

Acknowledgments

We acknowledge financial support from the Conselho Nacional de Desenvolvimento Cientifico e Tecnológico (CNPq) and the Fundação de Amparo à Ciência e Tecnologia do Estado de Pernambuco (FACEPE). The work was performed in the framework of the National Institute of Photonics (INCT de Fotônica) project and PRONEX/CNPq/FACEPE.

References and links

1. O. Wada, “Femtosecond all-optical devices for ultrafast communication and signal processing,” New J. Phys. 6, 183 (2004). [CrossRef]  

2. H. Lu, X. Liu, L. Wang, Y. Gong, and D. Mao, “Ultrafast all-optical switching in nanoplasmonic waveguide with Kerr nonlinear resonator,” Opt. Express 19(4), 2910–2915 (2011). [CrossRef]   [PubMed]  

3. H. M. Gibbs, Optical Bistability: Controlling Light with Light (Academic, 1985).

4. H. Ishikawa, Ultrafast All-Optical Signal Processing Devices (Wiley, 2008).

5. A. V. Kimel, “All-optical switching: three rules of design,” Nat. Mater. 13(3), 225–226 (2014). [CrossRef]   [PubMed]  

6. V. Lucarini, J. J. Saarinen, K.-E. Peiponen, and E. M. Vartiainen, Kramers-Kronig Relations in Optical Materials Research (Springer, 2005).

7. V. Mizrahi, K. W. Delong, G. I. Stegeman, M. A. Saifi, and M. J. Andrejco, “Two-photon absorption as a limitation to all-optical switching,” Opt. Lett. 14(20), 1140–1142 (1989). [CrossRef]   [PubMed]  

8. X. Hu, P. Jiang, C. Xin, H. Yang, and Q. Gong, “Nano-Ag: polymeric composite material for ultrafast photonic crystal all-optical switching,” Appl. Phys. Lett. 94(3), 031103 (2009). [CrossRef]  

9. Z. Li, X. Hua, Y. Zhang, Y. Fu, H. Yang, and Q. Gong, “All-optical switching via tunable coupling of nanocomposite photonic crystal microcavities,” Appl. Phys. Lett. 99(14), 141105 (2011). [CrossRef]  

10. C. B. de Araújo, T. R. Oliveira, E. L. Falcão-Filho, D. M. Silva, and L. R. P. Kassab, “Nonlinear optical properties of PbO-GeO2 films containing gold nanoparticles,” J. Lumin. 133, 180–183 (2013). [CrossRef]  

11. O. Sánchez-Dena, P. Mota-Santiago, L. Tamayo-Rivera, E. V. García-Ramírez, A. Crespo-Sosa, A. Oliver, and J.-A. Reyes-Esqueda, “Size-and shape-dependent nonlinear optical response of Au nanoparticles embedded in sapphire,” Opt. Mater. Express 4(1), 92–100 (2014). [CrossRef]  

12. P. Chakraborty, “Metal nanoclusters in glasses as non-linear photonic materials,” J. Mater. Sci. 33(9), 2235–2249 (1998). [CrossRef]  

13. J. Jayabalan, A. Singh, S. Khan, and R. Chari, “Third-order nonlinearity of metal nanoparticles: isolation of instantaneous and delayed contributions,” J. Appl. Phys. 112(10), 103524 (2012). [CrossRef]  

14. E. L. Falcão-Filho, C. B. de Araújo, A. Galembeck, M. M. Oliveira, and A. J. G. Zarbin, “Nonlinear susceptibility of colloids consisting of silver nanoparticles in carbon disulfide,” J. Opt. Soc. Am. B 22(11), 2444–2449 (2005). [CrossRef]  

15. E. L. Falcão-Filho, C. B. de Araújo, and J. J. Rodrigues Jr., “High-order nonlinearities of aqueous colloids containing silver nanoparticles,” J. Opt. Soc. Am. B 24(12), 2948–2956 (2007). [CrossRef]  

16. E. L. Falcão-Filho, R. Barbosa-Silva, R. G. Sobral-Filho, A. M. Brito-Silva, A. Galembeck, and C. B. de Araújo, “High-order nonlinearity of silica-gold nanoshells in chloroform at 1560 nm,” Opt. Express 18(21), 21636–21644 (2010). [PubMed]  

17. L. A. Gómez, C. B. de Araújo, A. M. Brito-Silva, and A. Galembeck, “Influence of stabilizing agents on the nonlinear susceptibility of silver nanoparticles,” J. Opt. Soc. Am. B 24(9), 2136–2140 (2007). [CrossRef]  

18. L. A. Gómez, C. B. de Araújo, A. M. Brito-Silva, and A. Galembeck, “Solvent effects on the linear and nonlinear optical response of silver nanoparticles,” Appl. Phys. B 92(1), 61–66 (2008). [CrossRef]  

19. Y. M. Wu, L. Gao, and Z. Y. Li, “The influence of particle shape on nonlinear optical properties of metal-dielectric composites,” Phys. Status Solidi, B Basic Res. 220(2), 997–1008 (2000). [CrossRef]  

20. R. Sato, M. Ohnuma, K. Oyoshi, and Y. Takeda, “Experimental investigation of nonlinear optical properties of Ag nanoparticles: Effects of size quantization,” Phys. Rev. B 90(12), 125417 (2014). [CrossRef]  

21. A. S. Reyna and C. B. de Araújo, “Nonlinearity management of photonic composites and observation of spatial-modulation instability due to quintic nonlinearity,” Phys. Rev. A 89(6), 063803 (2014). [CrossRef]  

22. A. S. Reyna and C. B. de Araújo, “Spatial phase modulation due to quintic and septic nonlinearities in metal colloids,” Opt. Express 22(19), 22456–22469 (2014). [PubMed]  

23. A. S. Reyna, K. C. Jorge, and C. B. de Araújo, “Two-dimensional solitons in a quintic-septimal medium,” Phys. Rev. A 90(6), 063835 (2014). [CrossRef]  

24. P. C. Lee and D. Meisel, “Adsorption and surface-enhanced Raman of dyes on silver and gold sols,” J. Phys. Chem. 86(17), 3391–3395 (1982). [CrossRef]  

25. A. M. Brito-Silva, L. A. Gómez, C. B. de Araújo, and A. Galembeck, “Laser ablated silver nanoparticles with nearly the same size in different carrier media,” J. Nanomater. 2010, 142897 (2010). [CrossRef]  

26. A. Takami, H. Kurita, and S. Koda, “Laser-induced size reduction of noble metal particles,” J. Phys. Chem. B 103(8), 1226–1232 (1999). [CrossRef]  

27. A. Pyatenko, M. Yamaguchi, and M. Suzuki, “Laser photolysis of silver colloid prepared by citric acid reduction method,” J. Phys. Chem. B 109(46), 21608–21611 (2005). [CrossRef]   [PubMed]  

28. M. Sheik-Bahae, A. A. Said, T.-H. Wei, D. J. Hagan, and E. W. Van Stryland, “Sensitive measurement of optical nonlinearities using a single beam,” IEEE J. Quantum Electron. 26(4), 760–769 (1990). [CrossRef]  

29. M. A. Duguay and J. W. Hansen, “An ultrafast light gate,” Appl. Phys. Lett. 15(6), 192–194 (1969). [CrossRef]  

30. H. Ma, A. S. L. Gomes, and C. B. de Araújo, “Measurements of nondegenerate optical nonlinearity using a two-color single beam method,” Appl. Phys. Lett. 59(21), 2666–2668 (1991). [CrossRef]  

31. P. N. Butcher and D. Cotter, The Elements of Nonlinear Optics (Cambridge University, 1990).

32. N. C. Kothari, “Effective-medium theory of a nonlinear composite medium using the T-matrix approach: Exact results for spherical grains,” Phys. Rev. A 41(8), 4486–4492 (1990). [CrossRef]   [PubMed]  

33. J. Jayabalan, “Origin and time dependence of higher-order nonlinearities in metal nanocomposites,” J. Opt. Soc. Am. B 28(10), 2448–2455 (2011). [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) Linear absorption spectra of the metal-colloid with f=4.0× 10 5 and acetone (cell thickness: 1 mm). (b) Size distribution histogram of the NPs after photofragmentation. A TEM image of the silver NPs is shown in the inset.
Fig. 2
Fig. 2 Normalized (a) closed-aperture and (b) open-aperture Z-scan profiles obtained for different NPs volume fractions. Laser peak intensity: 9.5 GW / c m 2 .
Fig. 3
Fig. 3 Dependence with the NPs volume fraction of: (a) NL refractive indices, (b) NL absorption coefficients, (c) total effective NL refractive index and (d) total effective NL absorption coefficient. In (a) and (b) the laser peak intensity was 9.5 GW / c m 2 . Normalized (e) Closed-aperture and (f) Open-aperture Z-scan profiles for f=5.9× 10 5 , obtained for different laser peak intensities. The solid lines were obtained from Eqs. (3) and (4).
Fig. 4
Fig. 4 (a) Transmitted Kerr signal in 532 nm for silver colloid for pump intensities of: I 1 =10 GW / c m 2 , I 2 =8.6 GW / c m 2 , I 3 =8.0 GW / c m 2 and I 4 =7.0 GW / c m 2 and I probe =0.1( I pump ) . The inset is the Kerr signal for CS2. (b) Dependence of | ΔT | / I pump as a function of the pump intensity. Volume fraction: 5.9× 10 5 .
Fig. 5
Fig. 5 Figures-of-merit for all-optical switching.

Tables (1)

Tables Icon

Table 1 NPs volume fraction, f, laser intensity, I, and the corresponding total NL refractive indices, nNL(I), total NL absorption coefficients, αNL(I), and figures-of-merit (T and W)

Equations (7)

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

Δ T PV I 0.396k L eff (1) n 2 +0.198k L eff (2) n 4 I+0.102k L eff (3) n 6 I 2 ,
ΔT I ( 2 ) 3 2 L eff (1) ( α 2 + α 4 I+ α 6 I 2 ).
T(z,Δ Φ 0 )1+ N=1 3 ( 4N )Δ Φ 0 (2N+1) z/ z 0 [ ( z/ z 0 ) 2 + (2N+1) 2 ] [ ( z/ z 0 ) 2 +1 ] N ,
T(z, q 0 ) 1 π q 0 ln[ 1+ q 0 exp( τ 2 ) ]dτ .
χ eff (3) =f L 2 | L | 2 χ np (3) + χ h (3) ,
χ eff (5) =f L 2 | L | 4 χ np (5) 6 10 f L 3 | L | 4 ( χ np (3) ) 2 3 10 fL | L | 6 | χ np (3) | 2 ,
χ eff (7) =f L 2 | L | 6 χ np (7) + 12 35 f L 4 | L | 6 ( χ np (3) ) 3 + 3 35 f | L | 8 [ 4 L 2 χ np (3) + | L | 2 ( χ np (3) ) * ] | χ np (3) | 2 4 7 fL | L | 6 [ 2 L 2 χ np (3) + | L | 2 ( χ np (3) ) * ] χ np (5) ,
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.