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

Third-order nonlinear optical properties of a multi-layer Al2O3/ZnO for nonlinear optical waveguides

Open Access Open Access

Abstract

This is a report of a study of the nonlinear optical properties of samples based on multiple Al2O3/ZnO bilayers fabricated by atomic layer deposition (ALD) in silica. The multi-layer configuration for samples consists of alternated layers of constant thickness of Al2O3x) and ZnOy) nanolaminates with a total thickness of ∼ 500 nm. The physical properties of the samples were characterized by means of TEM, spectrophotometry and variable angle spectroscopic ellipsometry. The absorptive and refractive contributions to the nonlinearity of the samples were studied by means of z-scan technique using a 100 fs at 800 nm. The nonlinear parameters, β and n2, are studied using different values of the layers thickness, Δx and Δy, in the nanolaminated stack. The possible applications in optical signal processing system are discussed by means of the figures of merit W and T.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In recent years there has been an increasing interest on the study of different materials for applications based in optical signal processing systems, which require a large nonlinearity and an ultra-fast response time [1, 2]. Different materials have been synthesized and studied for this purpose, among them nanocomposites with different geometries. Given the nanometric nature of some of the dimensions of the composites, their response to visible light can be considered as that of an effective medium. An interesting feature is the possibility of tailoring the nonlinear response of such effective media through manipulation of the physical properties [3], such as: composition, density, and the shape of the nanostructures [4].

The possibility of manipulating the optical properties of the nanostructured materials makes them of a great interest for applications in optical information processing systems [5]. Those applications include optical switches and other photonic devices with geometries based in waveguide form [2, 6]. Thereby, in order to achieve the integration of these nanomaterials for this purpose, fabrication of devices with precise control of their physical properties, represent a technological challenge. Since this is of interest for the optoelectronics and photonics industries [7,8], many methods for the production of nanostructured materials, either physical or chemical, have been implemented and evaluated [3,9–11].

Among the different techniques explored for producing nanostructures, the Atomic Layer Deposition (ALD) has proved to be a reliable technique for producing nanolayers with well controlled physical and chemical properties. In addition, the ALD technique [12], allows the obtention of ultrathin multilayer films composed of several bilayers with precise thickness control on nanometric scale [13]. The optical, electronic and physical properties of these materials will be a function of the layer thicknesses and growth temperature, as well as of the appropriate selection of materials [14,15]

Some studies of the third order nonlinear response of this kind of nanolayered materials have focused on the enhancement of χeff(3) of the composite through the manipulation of the effective dieletric constant eff, related to the effective refractive index neff=eff, of two ultrathin materials [16]. In this case, eff (using a E parallel to the plane of the layer) is given by 1/eff = fa/a + fb/b, with a and b the dielectric constants for materials a and b respectively, and fa and fb the corresponding volume fractions. For such material, the effective third-order nonlinear susceptibility can be written as:

χeff(3)=|effa|2(effa)2faχa(3)+|effb|2(effb)2fbχb(3).

Nevertheless, for E perpendicular to the plane of the layers, eff = faa + fbb and χeff(3) is given by [17]:

χeff(3)=faχa(3)+fbχb(3).

From here, it is easily seen that the nonlinear optical responses can be tuned through manipulation of the volume fraction of the nanolayers, the dielectric constants, and the interaction of E with the plane of layers.

On the other hand, there have been many studies of the optical response of ZnO, a good candidate for implementing different optical functions [11], including the fabrication of waveguides. This is due to its high linear refractive index n0, which results in a strong confining factor, their large nonlinear refractive index n2 and luminescence properties. Some studies [7] include the third-harmonic generation (THG) in ZnOAl2O3 bilayers synthesized by ALD, and its dependence on the layers widths. They found an increase of the THG in ZnO/Al2O3 nanolaminates through the control of the crystal structure of the ZnO without affecting the material volume [7].

In this work, we study the absorptive and refractive contributions to the nonlinear response, given by the nonlinear refractive index n2 (related to Re χ(3)(−ω; ω, −ω, ω)), and the two-photon absorption (TPA) coefficient β (related to Im χ(3)(−ω; ω, ω, −ω)), of samples based on multiple Al2O3/ZnO bilayers fabricated by ALD in silica. The absorptive and refractive contributions to the nonlinearity of the samples were studied by means of the z-scan technique in the near-infrared regime. A study of the nonlinear parameters β and n2 is made for different Al2O3/ZnO ratios in the nanolaminated stack. Finally, we discuss the possible applications of this materials in optical signal processing by means of figures of merit W and T.

2. Sample preparation and linear optical properties

Ultrathin multilayer films based on Al2O3/ZnO bilayer stacks were deposited on Si (100) and SiO2 substrates in order to study their linear and nonlinear optical properties. We fabricated the samples by ALD technique using a Beneq TFS 200 ALD viscous-flow reactor at 200 °C. Trimethylaluminum (TMA) (Strem 93 − 1360) and Diethylzinc (DEZ) (Strem 93 − 3030) were used as the aluminum and zinc sources, respectively. For this process, deionized water was used as the oxidant agent, all the precursor were used at room temperature. UHP N2 (< 106 ppm O2) was used as carrier/purging gas purified by means of a Centorr 2G − 100 − SS − 120 gettering furnace.

The scheme in Fig. 1(a) depicts the pulse sequence of ALD cycles used to grow the Al2O3/ZnO ultrathin multilayer stacks. The complete ALD cycle for both precursors with the exposure dose and purge time, respectively, is also shown in Fig. 1(a). In the scheme, the parameter d corresponds to the number of repetitions in the recipe, necessary to obtain the nominal total programmed thickness (∼ 500 nm). In this arrangement, the Al2O3 layers have a thickness Δx (nm), while Δy (nm) corresponds to the ZnO layers. Both, Δx and Δy, were varied for different samples.

 figure: Fig. 1

Fig. 1 a) The sequence of ALD cycles for multi-layer fabrication samples, and b) the geometry of the resulting materials.

Download Full Size | PDF

A set of 5 samples, labeled MΔxy, with different number of cycles were prepared. The total thickness of the samples was kept at ∼ 500 nm in order to keep their absorption coefficients values sufficiently low to get a useful throughput and to avoid any damage to the samples. Additionally, this thickness value allows efficient light coupling when the samples are used to produce slab or channel waveguides. Figure 1(b) shows the multilayer scheme of the bilayers MΔxy, deposited in SiO2 substrate for the nonlinear measurements.

Figure 2 shows a cross-sectional TEM image of a Al2O3/ZnO nanolaminate, in this case (Al2O3/ZnO)5/SiO2/Si, in which Al2O3 and ZnO layers were made by 100 and 50 ALD cycles, respectively. At the bottom of Fig. 2(b) the Si wafer is clearly observed, as well as a native SiO2 thin layer. As is demonstrated on Fig. 2(a), well defined nanolaminated structures have been grown, composed by gray layers (Al2O3) and darker ones (ZnO) with very flat interfaces between the different metal oxides layers. The single layer thickness obtained was approximately 10 nm for Al2O3 and 10 nm for ZnO, in agreement with the growth rate per cycle calculated for both Al2O3 and ZnO. It is easy to discern that the Al2O3 layer is amorphous, while the ZnO layer presents crystalline domains (Fig. 2(b)). The above result indicates that the ALD method, has a high accuracy in the thickness control when the multilayers are grown on flat substrates.

 figure: Fig. 2

Fig. 2 Nanolaminated were observed by TEM using a JEOL JEM2010 microscope at 200 kV. Sample was prepared through focus ion beam (FIB) on a JEOL FIB-4500 SEM.

Download Full Size | PDF

2.1. Linear characterization

2.1.1. Absorption spectra

In order to characterize the linear optical properties of the resulting samples, the optical density spectra of the materials were taken using a Hitachi 7000 spectrophotometer. Figure 3 shows the optical density spectra for all samples MΔxy. Also the Fig. 3 shows a high transparency in the visible region of all samples produced, and periodic oscillations in the curves, due to interference effects related to the thickness and refractive index of the stack. In addition, the samples show the typical absorption edge towards the UV, that starts below 400 nm, for thin films composites containing ZnO.

 figure: Fig. 3

Fig. 3 Absorbance spectra of samples MΔxy. The values of Δx and Δy are in nm.

Download Full Size | PDF

2.1.2. Ellipsometry measurements

It is important to measure the linear refractive index in order to observe the dependence of eff with the manipulation of the amount of Al2O3 and ZnO and its correlation with the nonlinear optical parameters of the materials. Briefly, mixing 2 different materials in a nanolaminate structure will produce intermediate refractive index values that will go from 1.65, for Al2O3, to 2 for ZnO (at 632 nm). To calculate this, we use the variable angle spectroscopic ellipsometry technique that can give us information of the total multilayer stack thickness and its refractive index across the optical spectrum [8]. Hence, we use a J. A. Woollam M − 2000 ellipsometer to measure the amplitude and phase difference between reflection coefficients (rp and rs) from λ = 350 nm to 1000 nm at different angles (θ = 45°, 55°, 65° and 75°). With the results obtained, we use the Tauc-Lorentz model in order to get a fit, and extract the refractive index for all samples.

The dispersion curves of the samples fabricated on Si are shown in Fig. 4. The refractive index curves of samples M1,Δyx = 1 nm and Δy is varying), are presented in Fig. 4(a), while Fig. 4(b), exhibits the refraction index curves of samples MΔx,10x is varying and Δy = 10 nm is fixed). The refractive index of the samples shown in Fig. 4(a) exhibit an abrupt change going from samples M1,2 to samples M1,5 and M1,10. This is probably due to the significant change in ZnO content. A quasi-linear control of the refractive index could be observed in Fig. 4(b) due to the changing amount of Al2O3, as expected. These results are consistent with those previously reported in the literature, where a modulation of the linear refraction index has been demonstrated in samples consisting of Al2O3/ZnO alternating layers of different thickness [7, 18], having overall values of ∼ 500 nm and ∼ 100 nm, respectively.

 figure: Fig. 4

Fig. 4 Refractive index for a) M1,Δy and b) MΔx,10 samples.

Download Full Size | PDF

The total thickness, amount of the materials deposited on the substrate, the aspect ratio and the number of interfaces are shown in the Table 1.

Tables Icon

Table 1. The total thickness, amount of the materials deposited on the substrate, the aspect ratio and the number of interfaces of the multilayer MΔxy samples.

3. Nonlinear optical experiments

3.1. Experimental setup

The nonlinear response of the samples was studied using the z-scan technique with fs pulses. This technique allows the resolution of the refractive and absorptive contributions to the nonlinear response, as well as to the determination of the sign of both contributions [19]. Briefly, in this technique the nonlinear sample is scanned across the focal plane of a lens, where a spatially Gaussian pulse is tightly focused. In this sense, the transmittance through an aperture located in the far field is measured as the nonlinear sample is scanned across the focal plane of the lens. Hence, the apertured detector, is sensitive to changes in both nonlinear absorption and refraction (closed aperture z-scan). Nevertheless, we can measure the nonlinear absorption through a slightly different open-aperture configuration, where the whole transmitted beam is detected (open-aperture Z-scan), which then will be sensitive to only nonlinear absorption. Thereby, is possible to discern the nonlinear optical responses by a simple division of the closed-aperture by the open-aperture transmittance traces [20].

In the setup employed, illustrated in Fig. 5, the light source is a 76 MHz repetition rate Ti:sapphire laser with 100 fs duration pulses at 800 nm wavelength. The incident laser power was controlled by a half-wave plate followed by a polarizer prism and monitored by detector D1. The sample was mounted in a stage that moves along the beam propagation direction (z axis). The beam waist at z = 0 was 47 μm. Pre-focal and post-focal positions correspond to z < 0 and z > 0, respectively. After crossing the sample, the laser beam passes through a circular aperture, placed in the far-field region, being detected by detector D2.

 figure: Fig. 5

Fig. 5 Z-scan setup employed in the experiments,BS is a beam splitter, L1 is a lens with focal length of 20 cm, D1 is the detector whose monitors the input laser power, while the D2 is de detector to monitoring the closed aperture z-scan signal, respectively.

Download Full Size | PDF

The experimental data for the closed to open aperture ratio can be used to extract the nonlinear refractive index n2 by measuring the peak to valley transmittance change ΔTpv, using ΔTpv = 0.406kLeffn2I0, where k = 2π/λ, Leff = [1 − exp (−α0L)]/α0, L is the sample length, α0 is the linear absorption coefficient and I0 is the laser intensity at the focal plane. In the same way, the open aperture z-scan data (for S = 1), can be used to get the two-photon absorption coefficient β, through T(z=0)=1βI0Leff/8 [21].

3.2. Results and discussion

Figure 6 displays experimental open and closed-aperture z-scan results for two representative samples: M1,2 and M1,5, the circles represent the experimental data and the continuous lines are the theoretical fits to them. Figures 6(a) and 6(b) show the open and closed z-scan data for M1,2, while Fig. 6(c) and Fig. 6(d) the open and closed z-scan data for M1,5, respectively. Notice that in Fig. 6(a), there is not discernible nonlinear absorption, since there is no changes in the normalized transmittance. On the other hand, Fig. 6(c), shows the typical two photon absorption profile, where a minimum is discernible in the focal plane. For the closed-aperture z-scan results in Fig. 6(b) and Fig. 6(a), it is easy to discern the typical signature of a positive n2, a pre-focal minimum followed by a post-focal maximum. The irradiance used for all these measurements was varied from 75.8 to 341 MW/cm2 (see Table 2).

 figure: Fig. 6

Fig. 6 Open and closed-aperture z-scan results for the samples M1,2 a) open, b) closed; and M1,5 c) open, and d) closed. For these experiments, the irradiances used were I0 = 152 MW /cm2 and I0 = 341 MW /cm2, respectively.

Download Full Size | PDF

The nonlinear parameters β and n2 for all samples are displayed in the Fig. 7. These values have been obtained by fitting the theory, according to [19, 21], to the experimental open and closed z-scan data. The black points in Fig. 7(a), are the values for the two photon absorption coefficient β for each sample, while the black points in Fig. 7(b), are the n2 obtained values for each sample.

 figure: Fig. 7

Fig. 7 Correlation between the linear and nonlinear optical parameters. The Fig (a) show the linear absorption α at 2ω and the nonlinear optical absorption coefficient β. The Fig (b) show the linear optical index n0 at ω and the nonlinear refractive index n2.

Download Full Size | PDF

The nonlinear optical parameters have been obtained far away of resonance, nevertheless the linear optical absorption is slightly higher at the two photon regime, i.e. at twice the laser frequency, so that TPA process is possible. This is confirmed by the open z-scan data obtained, where the two photon absorption effect is observed (Fig. 6). Due to this, we have added the value of α(2ω) in Fig. 7(a), in order to check if there is a correlation between them. The results indeed suggest the small differences in linear absorption at 2ω generate a change in the non-linear absorption at ω in the samples studied, as expected. In the same way we have plotted in Fig. 7(b) the data of the linear refractive index n0(ω) together with the measured n2 values. The results show a similar behaviour, that there is a correlation between the linear and nonlinear refraction coefficients n0 and n2, thereby to eff for both samples as well. Even though in the z-scan geometry employed, the electric field vector is parallel to plane of the layers, we still expect an enhancement in the eff and thereby in n2 as well. If the multilayers were to be used as waveguides we could expect in principle a further enhancement in χeff(3).

Comparing our results with the literature, the nonlinear refractive index for Al2O3 bulk has been calculated and its value is around n2 = 1.33 × 10−9 cm2/GW [22], and on the other hand Lin et al [23] have obtained a n2 = 2.65 × 10−2 cm2 /GW value for ZnO thin films samples. Both values of the nonlinear refractive index are lower than the one we have determined for the M1,10 bilayer sample.

Regarding the physics behind the nonlinear response, intra-pulse thermal effects are effectively discarded, due to to relatively long rise time of the temperature change induced by a single fs pulse, given by τrise = ω0/vs, where ω0 is the beam waist and vs is the sound velocity in the sample. However, the high repetition rate of the pulses (76MHz) could in principle imply the possibility of pulse to pulse thermal effects [24–26], but, given the off-resonance characteristics of the interaction, heating by linear absorption can be effectively ruled out.

In order to evaluate the potential of the material for application in all-optical switching devices, it is useful to calculate the figures of merit devised to assess the potential of different materials [27]. The first figure of merit W is given by:

W=Δnmaxλα0,
where Δnmax is the maximum achievable refractive index change (when saturable absorption is present, Δnmax = n2Isat, with Isat the saturation irradiance). This figure of merit indicates whether a given refractive index change can be achieved within a linear absorption length (Labs = 1/α0). The second one, T given by:
T=βλn2,
indicates whether nonlinear absorption can limit or not the applicability of the materials; suggesting in this case that a given phase change can be obtained within a nonlinear absorption length LNL = 1/(βI). The acceptable values of W and T are device dependent, but we can safely assume that W > 1, and T < 1.

In our case, we use the experimental data taken at the highest irradiance employed (where I0 < Isat) in order to estimate Δnmax, and hence a lower bound for W. Although we were not able to measure the actual β value for M1,2, we can calculate the smaller β value than our apparatus can detect in order to give a possible value to T. To the case of the other samples, we use the values of n2 and β obtained, to calculate in turn upper bounds for the values of W and T. The calculated W and T values are shown in Table 2.

At this point, we have obtained the limit values for the W and T figures of merit. Although we were not able to reach the Isat (which could cause an enhancement in W value through Δnmax), the W values obtained from all samples are close to 1, even could be higher to this value. In the case of the T values, the β coefficient calculated for the sample M1,2 suggest a value closer to 1, since the β coefficient could not be obtained with the irradiances used, while for M1,10 sample, the T value is marginally close to 1 with β coefficient obtained.

Tables Icon

Table 2. Nonlinear optical parameters n2 and β, and figures of merit W and T, evaluated under the irradiances showed for samples MΔxy.

4. Conclusions

In summary, this work reports the fabrication by ALD of Al2O3/ZnO multilayer samples. The nonlinear optical parameters n2 and β were evaluated for the samples with differences in the thickness of the Al2O3 and ZnO layers. The samples present a large positive nonlinear refractive index, that can be manipulated through the fabrication parameters. Comparing the results obtained, we note that to the samples with a higher content of ZnO with respect to Al2O3 amount, shows an enhancement in the nonlinear refractive index measured. Although we do not seem to reach irradiance values close to a saturation irradiance Isat, and considering the large n2 and low β values obtained, we claim that composites with Al2O3/ZnO multilayer could have a good potential for application in all-optical signal processing devices.

Funding

Dirección General de Asuntos del Personal Académico DGAPA-UNAM (PAPIIT: IN113219, IN110018, IN112117, IA101018); Dirección General de Asuntos del Personal Académico DGAPA-UNAM (PAPIME: PE 100318, PE101317); FORDECYT-CONACYT 272894; CONACYT-Mexico 222485.

Acknowledgments

The authors would like to thank valuable technical support to David Dominguez, Eloisa Aparicio, Eduardo Murillo, Israel Gradilla, Francisco Ruiz and Jaime Mendoza. Also BCU acknowledges the Postdoctoral Scholarship provided by DGAPA-UNAM.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. U. Kreibig and M. Vollmer, “Optical Properties of Metal Clusters,” (Springer, 1995). [CrossRef]  

2. B. Can-Uc, R. Rangel-Rojo, H. Márquez, L. Rodríguez-Fernández, and A. Oliver, “Nanoparticle containing channel waveguides produced by a multi-energy masked ion-implantation process,” Opt. Express 23, 3176–3185 (2015). [CrossRef]   [PubMed]  

3. R. Rangel-Rojo, J. McCarthy, H. T. Bookey, A. K. Kar, L. Rodríguez-Fernández, J. C. Cheang-Wong, A. C. Sosa, A. López-Suares, A. Oliver, V. Rodríguez-Iglesias, and H. G. Silva-Pereyra, “Anisotropy in the nonlinear absorption of elongated silver nanoparticles in silica, probed by femtosecond pulses,” Opt. Commun. 282, 1909–1912 (2009). [CrossRef]  

4. K. L. Kelly, E. Coronado, L. L. Zhao, and G. C. Schatz, “The optical properties of metal nanoparticles: the influence of size, shape, and dielectric environment,” J. Phys. Chem. B 107, 668–677 (2003). [CrossRef]  

5. C. Torres-Torres, J. Bornacelli, B. Can-Uc, H. G. Silva-Pereyra, L. Rodríguez-Fernández, M. Avalos-Borja, G. J. Labrada-Delgado, J. C. Cheang-Wong, R. Rangel-Rojo, and A. Oliver, “Coexistence of two-photon absorption and saturable absorption in ion-implanted platinum nanoparticles in silica plates,” J. Opt. Soc. Am. B 35, 1295–1300 (2018). [CrossRef]  

6. B. E. Olsson, L. Rau, and D. J. Blumenthal, “Wdm to otdm multiplexing using an ultrafast all-optical wavelength converter,” IEEE Photonics Technol. Lett. 13, 1005–1007 (2001). [CrossRef]  

7. L. Karvonen, A. Säynätjoki, Y. Chen, H. Jussila, J. Rönn, M. Ruoho, T. Alasaarela, S. Kujala, R. A. Norwood, N. Peyghambarian, K. Kieu, and S. Honkanen, “Enhancement of the third-order optical nonlinearity in zno/al2o3 nanolaminates fabricated by atomic layer deposition,” Appl. Phys. Lett. 103, 031903 (2013). [CrossRef]  

8. J. López, H. A. Borbón-Nunez, E. G. Lizarraga-Medina, E. Murillo, R. Machorro, N. Nedev, H. Marquez, M. H. Farías, H. Tiznado, and G. Soto, “al2o3 − y2o3 ultrathin multilayer stacks grown by atomic layer deposition as perspective for optical waveguides applications,” Opt. Mater. 72, 788–794 (2017). [CrossRef]  

9. A. Oliver, J. A. Reyes-Esqueda, J. C. Cheang-Wong, C. E. Román-Velasquez, A. Crespo-Sosa, L. Rodriguez-Fernandez, J. A. Seman, and C. Noguez, “Controlled anisotropic deformation of ag nanoparticles by si ion irradiation,” Phys. Rev. B 74, 245425 (2006). [CrossRef]  

10. N. D. Fatti and F. Vallee, “Ultrafast optical nonlinear properties of metal nanoparticles,” Appl Phys B 73, 383–390 (2001). [CrossRef]  

11. C. Torres-Torres, B. A. Can-Uc, R. Rangel-Rojo, L. Castaneda, R. Torres-Martínez, C. I. García-Gil, and A. V. Khomenko, “Optical kerr phase shift in a nanostructured nickel-doped zinc oxide thin solid film,” Opt. Express 21, 21357–21364 (2013). [CrossRef]   [PubMed]  

12. K. Hyungjun, L. Han-Bo-Ram, and W. J. Maeng, “Applications of atomic layer deposition to nanofabrication and emerging nanodevices,” Thin Solid Films 517, 2563–2580 (2009). [CrossRef]  

13. S. M. George, “Atomic layer deposition: An overview,” Chem. Rev. 110, 111–131 (2010). [CrossRef]  

14. D. F. Schmidt, “Nanolaminates–bioinspired and beyond,” Mater. Lett. 108, 328–335 (2013). [CrossRef]  

15. R. Viter, I. Baleviciute, A. A. Chaaya, L. Mikoliunaite, Z. Balevicius, A. Ramanavicius, A. Zalesska, V. Vatamana, V. Smyntyna, Z. Gertnere, D. Erts, P. Miele, and M. Bechelanye, “Optical properties of ultrathin al2o3/zno nanolaminates,” Thin Solid Films 594, 96–100 (2015). [CrossRef]  

16. G. L. Fischer, R. W. Boyd, R. J. Gehr, S. A. Jenekhe, J. A. Osaheni, J. E. Sipe, and L. A. Weller-Brophy, “Enhanced nonlinear optical response of composite materials,” Phys. Rev. Lett. 74, 1871–1874 (1994). [CrossRef]  

17. R. W. Boyd and J. E. Sipe, “Nonlinear optical susceptibilities of layered composite materials,” J. Opt. Soc. Am. B 11, 297–303 (1994). [CrossRef]  

18. J. López, E. Solorio, H. A. Borbón-Nunez, F. F. Castillón, R. Machorro, N. Nedev, M. H. Farías, and H. Tiznado, “Refractive index and bandgap variation in al2o3-zno ultrathin multilayers prepared by atomic layer deposition,” J. Alloys Compd. 691, 308–315 (2017). [CrossRef]  

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

20. M. G. Kuzyk and C. W. Dirk, Characterization techniques and tabulations for organic nonlinear materials (CRC Press, 1998.

21. P. Poornesh, G. Umesh, P. K. Hegde, M. G. Manjunatha, K. B. Manjunatha, and A. V. Adhikari, “Studies on third-order nonlinear optical properties and reverse saturable absorption in polythiphene (methylmethacrylate) composites,” Appl. Phys. B 97, 117–124 (2009). [CrossRef]  

22. B. Can-Uc, R. Rangel-Rojo, A. Pena-Ramírez, C. B. de Araújo, H. T. M. C. M. Baltar, A. Crespo-Sosa, M. L. Garcia-Betancourt, and A. Oliver, “Nonlinear optical response of platinum nanoparticles and platinum ions embedded in sapphire,” Opt. Express 24, 9955–9965 (2016). [CrossRef]   [PubMed]  

23. J. H. Lin, Y. J. Chen, H. Y. Lin, and W. F. Hsieh, “Two-photon resonance assisted huge nonlinear refraction and absorption in zno thin films,” J. Appl. Phys. 97, 033526 (2005). [CrossRef]  

24. M. Falconieri, “Thermo-optical effects in z-scan measurements using high-repetition-rate lasers,” J. Opt. A: Pure Appl. Opt. 1, 662–667 (1999). [CrossRef]  

25. A. Gnoli, L. Razzari, and M. Righini, “Z-scan measurements using high repetition rate lasers: how to manage thermal effects,” Opt. Express 13, 7976–7981 (2005). [CrossRef]   [PubMed]  

26. B. Can-Uc, R. Rangel-Rojo, L. Rodríguez-Fernández, and A. Oliver, “Polarization selectable nonlinearities in elongated silver nanoparticles embedded in silica,” Opt. Mater. Express 3, 2012–2021 (2013). [CrossRef]  

27. G. I. Stegeman, “All-optical devices: materials requirements,” Proc. SPIE 1852, 75–89 (1993). [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 (7)

Fig. 1
Fig. 1 a) The sequence of ALD cycles for multi-layer fabrication samples, and b) the geometry of the resulting materials.
Fig. 2
Fig. 2 Nanolaminated were observed by TEM using a JEOL JEM2010 microscope at 200 kV. Sample was prepared through focus ion beam (FIB) on a JEOL FIB-4500 SEM.
Fig. 3
Fig. 3 Absorbance spectra of samples MΔxy. The values of Δx and Δy are in nm.
Fig. 4
Fig. 4 Refractive index for a) M1,Δy and b) MΔx,10 samples.
Fig. 5
Fig. 5 Z-scan setup employed in the experiments,BS is a beam splitter, L1 is a lens with focal length of 20 cm, D1 is the detector whose monitors the input laser power, while the D2 is de detector to monitoring the closed aperture z-scan signal, respectively.
Fig. 6
Fig. 6 Open and closed-aperture z-scan results for the samples M1,2 a) open, b) closed; and M1,5 c) open, and d) closed. For these experiments, the irradiances used were I0 = 152 MW /cm2 and I0 = 341 MW /cm2, respectively.
Fig. 7
Fig. 7 Correlation between the linear and nonlinear optical parameters. The Fig (a) show the linear absorption α at 2ω and the nonlinear optical absorption coefficient β. The Fig (b) show the linear optical index n0 at ω and the nonlinear refractive index n2.

Tables (2)

Tables Icon

Table 1 The total thickness, amount of the materials deposited on the substrate, the aspect ratio and the number of interfaces of the multilayer MΔxy samples.

Tables Icon

Table 2 Nonlinear optical parameters n2 and β, and figures of merit W and T, evaluated under the irradiances showed for samples MΔxy.

Equations (4)

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

χ eff ( 3 ) = | eff a | 2 ( eff a ) 2 f a χ a ( 3 ) + | eff b | 2 ( eff b ) 2 f b χ b ( 3 ) .
χ eff ( 3 ) = f a χ a ( 3 ) + f b χ b ( 3 ) .
W = Δ n max λ α 0 ,
T = β λ n 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.