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

Large-area bulk self-assembly of plasmonic nanorods in nematic liquid crystal via surface-mediated alignment

Open Access Open Access

Abstract

In this paper, we introduce a novel method to organize gold nanorods (GNRs) on a large scale by means of nematic lyotropic liquid crystal (LC) and surface-induced bulk alignment. LC is unidirectionally aligned via surface-treatment, and then gets GNRs organized with a negative order parameter in bulk through elastic force. The self-assembled GNRs (with a considerable high concentration of about 6 wt%) in LC exhibit a strong polarization-dependent localized surface plasmon resonance (LSPR). Polarized optical microscopic imaging, polarization-dependent extinction and scattering spectra of LSPR are utilized to characterize the spatial orientation of GNRs. This method is compatible with the state-of-the-art LC display technique and allows for large-scale fabrication of optical metamaterials, plasmonic polarizers comprised of self-aligned anisometric nanoparticles with extremely low cost.

© 2013 Optical Society of America

1. Introduction

The self-assembly of metallic nanomaterials into functional superstructures in two- or three-dimensional space by way of the “bottom-up” method is of importance for the nano-fabrication technology of optical devices, such as polarization components, information storage devices [1] and optical metamaterials [2,3]. Several methods have been used to assemble nanoparticles into different scales, such as polymer matrix [4,5], nanostructured surface [6], and DNA conjugation [7]. Liquid crystal (LC) is a fascinating medium that can be utilized to assemble the micro- or nanoparticles into pre-engineered structures by controlling the assembly in one or more dimensions through combining a great variety of mesophase morphologies and giving easy responses to external stimulated field or boundary conditions [811]. Different nanostructured materials of a variety of shapes such as metallic nanoparticles [8,12], semiconductor nanorods [13], and carbon nanotubes [14] have been organized into structures to exhibit specific nano- and micro-scale arrangements in their LC host by taking advantage of the orientational order of LCs. Recently, the self-assembly of gold nanorods (GNRs) in nematic, hexagonal and cholesteric mesophases has been demonstrated, and the alignment of GNRs could be achieved by shear stress and magnetic fields [15,16]. However, these methods are limited to either low order parameter of GNRs or complex facilities. A more general method for aligning GNRs on a large scale remains a challenge.

Here we demonstrate that the colloidal GNRs are designed to orient in a pre-defined order on a large scale embedded in a LC matrix. The LC host is aligned via the surface-mediated force of rubbing the surface, which then forces GNRs to align perpendicular to the directors in the bulk. This aligned hybrid GNRs-LC system shows strong polarization-dependent extinction and scattering on a large scale. We characterize the alignment of GNRs and their optical properties by polarized optical microscopy, dark-field microscopy and polarized extinction/scattering spectra. This technique for organizing nanoparticles is compatible with the state-of-the-art LC display technology and is of potential importance in fabrication of large-area optical devices via self-assembled nanomaterials in information display, optical metamaterials and polarized plasmonic components.

2. Experiment

The GNR-LC composites are constituted of GNRs in a Nc phase. GNRs with a mean width of 20 nm and a mean length of 50 nm were synthesized using a seed-mediated method [17]. The concentration of GNRs is 0.3 nM given by the inductively coupled plasma mass spectrometry (ICP-MS) measurement. Thiol-terminated methoxy-poly(ethylene glycol) (mPEG-SH, JemKem Technology) was used to functionalize GNRs in order to stabilize them [18]. First, the dispersion of GNRs was centrifuged at 9000 rpm (corresponding to 7516 × g) for 20 min and then resuspended to 1 mL of deionized water. Then, 250 μL of an aqueous solution with 2 mM of 5 kDa mPEG-SH were added into 5 mL of a 3.5 nM GNRs dispersion. The mixture was left overnight and then purified via centrifugation to eliminate the excess mPEG-SH. The nematic lyomesophase was made by the sodium decylsulfate/decanol/water (SDS/DeOH/H2O) ternary system according to the phase diagram in [19] and [20]. Briefly, the hybrid GNRs-LC composites were prepared by adding 38.2 wt% SDS, 7 wt% DeOH and 54.8 wt% GNRs aqueous solution to a centrifuge tube successively and mixing them quickly. The mixture was under sonification and centrifuge several times to unify the compounds. The final concentration of GNRs is about 6 wt % (3 × 10−7 M) calculated by the total weight of GNRs-LC composites and the original concentration of mPEG-GNRs. The GNRs-LC system shows hexagonal and nematic phases (calamitic, discotic and biaxial) when the temperature varies and the phases are confirmed by their optical textures. Several experiments have shown that the building block of this ternary lyomesophase is biaxial micelle [2123]. The phases differ from each other by orientational fluctuations around different axes of symmetry of these micelles.

The surface-imposed alignment of lyotropic LC in Nc phase is based on the fact that a rubbing surface can induce a preferential orientation of the micelles [24,25]. The surface treatment creates an easy axis for biaxial micelles, and the far-field director N0 of LC is along the rubbing direction. We spin-coated a polymethyl methacrylate (PMMA) film 500nm thick on a glass plate. The unidirectional alignment of LC was achieved by rubbing the PMMA film with soft velvet or lens paper 20 times. The LC cell was made by two pieces of glass which were covered with such alignment layers in the inner sides. The LC cell was 5-20μm thick separated by SiO2 spacers. The GNRs-LC composite was then injected into the LC cell and the cell was then sealed by UV glue.

Polarized optical microscopy (POM) and polarized dark-field microscopy were used to characterize the alignment of LC and GNRs, respectively. An Olympus BX-51 mounted with a CCD camera (company) with 10 × , 20 × , 50 × and 100 × (oil immersion) objectives were used. A microscope mounted with a microspectrometer (USB2000, Ocean) was used to measure the polarization-dependent extinction spectra. The experimental setup for the transmittance measurements consists of an incident light from a halogen tungsten lamp, crossing a bright-field condenser, the GNRs-LC sample, a polarizer and finally a spectrometer. The scattering spectra were measured with a similar method but with a dark-field condenser instead of a bright-field condenser.

3. Results and discussion

The uniform alignment of calamitic nematic (Nc) lyotropic LC phase by surface-induced force was characterized by Polarized optical microscopy (POM) observation. The texture of LC appears to be totally dark when the rubbing direction k is parallel or perpendicular to the polarization of incident light P and brightest when k is oriented at 45°with respect to P, in Figs. 1(a) and 1(b). The orientation of the director of LC (N) is confirmed by inserting a phase compensator (530nm λ−plate) into the POM microscope (Fig. 1(c)). Because the Nc phase is optically negative (ne<no) [20], the blue interference color shows that N is along k in the laboratory coordination, in Fig. 1(c). Thus, in the presence of micro-grooves by rubbing the surface, N is parallel to k, while without any treatment of the surface layer N only tends to orient in the plane parallel to the substrate surface under the influence of the surface.

 figure: Fig. 1

Fig. 1 (a, b) Polarized optical microscopy images of GNR-doped LC for two different rubbing directions (indicated by k) between crossed polarizers (indicated by (P) and (A), showing high-quality long-range alignment. (c) Polarized optical microscopy images of GNR-doped LC with a phase compensator (530nm λ−plate, slow axis γ) inserted to the POM, showing the alignment of nematic LC is along the rubbing direction. The scale bars are 50μm.

Download Full Size | PDF

The orientation of Nc near the surface is elastically transferred to the bulk of the Nc phase and the whole LC sample is oriented at N along k. Because of the alignment of GNRs in the LC matrix, the collective surface plasmon resonance (SPR) is excited and the sample shows strong polarization and wavelength dependent extinction. These extinction spectra were measured by micro-spectrometer mounted in the microscope with a analyzer (see Fig. 2(b)). The GNRs-LC composite shows a red color when PN (or k), while it becomes green when Pk, in Figs. 2(c)-2(d). By means of this method, GNRs in LC can be aligned in a device-scale (on the order of inches) and is shown in the insets of Figs. 2(c) and 2(d). Based on localized SPR theory of nanorods, when the polarization of incident light is parallel to the long axis of a nanorod, the longitudinal SPR mode is excited and the nanorod absorbs and scatters light strongly, while the transverse SPR mode is excited when the polarization of incident light is perpendicular to the long axis of nanorod. Surprisingly, the long axis of GNRs in the Nc matrix is perpendicular to N (or k) in the bulk. The optical observation along with rotational symmetry reveal that GNRs are oriented in the plane perpendicularly to N there is no preferred orientation of GNRs in the plane, as illustrated in Fig. 2(a). This alignment is also observed in the cholesteric phase of nematic lyomesophase and can be explained by the phenomenological model that has been established [16]. This alignment can be understood based on the absorption and anchoring of rod-like building-block of LC by the flexible polymer chains capped on GNRs [26]. The nematic lyotropic LC is composed of biaxial micelles and represented by three symmetry axes. When the micelles form Nc phase, the orthorhombic micelles fluctuate around their longest axis and degenerate to rod-like shape [2123]. The flexible chains of mPEG radiate from the surface of GNRs in the aqueous solution because water is good solvent for mPEG. Then the rod-like micelles will absorb and insert into this polymer chains, which leads to the rod-like micelles orienting perpendicular to the surface of GNRs. Because there are more polymer chains on the side face of GNRs than the ends, this results in the long axis of GNRs orienting perpendicular to the director of Nc. Combining with the flexibility of the polymer chains, and the minimization of the anchoring energy of micelles on the GNRs surface along with the uniform alignment of the far field director N0 lead to the long axis of GNRs being perpendicular to N0 (or k).

 figure: Fig. 2

Fig. 2 (a) Schematic illustration of GNR alignment with a negative order parameter in the nematic LC. Red lines indicate the uniform director field of LC host. (b) The experiment setup used to measure the polarization-dependent extinction spectra of aligned GNR-LC composites. (c,d) The transmission optical microscopy images obtained for unpolarized light passing first through the cell with aligned GNRs and then through a linear polarizer (P) either perpendicular (c) or parallel (d) to the rubbing direction k. The insets of (c) and (d) show the images of the whole sample (in a centimeter scale) in two orthogonal cases. The scale bars are 50μm.

Download Full Size | PDF

When the polarization of incident light is in the intermediate angles between the long and short axes of GNRs, the aligned GNRs show angular dependence of extinction, in Fig. 3(a). The intensity of extinction varies from the minimum when Pk to the maximum when Pk. The imperfect suppression of the longitudinal peak when Pk derives from the limited order parameter of GNRs in Nc. The angular dependent extinction coefficient is shown in Fig. 3(b). The scalar order parameter of GNRs is defined as SGNR=P2(cosθ), where P2(cosθ) is the second Legendre polynomial and θ is an angle between a nanorod and N. In this case GNRs show a negative order parameter with respect to N. The alignment of GNRs could be characterized by its scalar order parameter, which is calculated as , where αext and αext are the extinction coefficients of the GNRs-LC composites when Pk and Pk, respectively [27]. By using and at the wavelength of the longitudinal SPR peak, we obtain SGNR = −0.35 ± 0.01, while the order parameter of GNRs with a perfect negative order is −0.5. The order parameter of GNRs could be further improved to approach −0.5 by decreasing the thickness of LC cell or using GNRs with a larger aspect ratio.

 figure: Fig. 3

Fig. 3 (a) The measured polarization-sensitive extinction spectra of GNRs for different angles between the linear polarizer and the rubbing direction. (b) The measured extinction value at the longitudinal SPR peak as a function of the angle between the linear polarizer and the rubbing direction.

Download Full Size | PDF

The dark-field microscopy possesses the ability to directly visualize the orientation of single GNR by detecting the polarization-dependent scattering of GNR. The intensity of scattering of GNRs is maximized when the polarization of incident light is parallel to the long axis of GNRs, while it becomes minimal when the polarization of incident light is parallel to the short axis of GNR. In the dark-field microscopy observation of aligned GNRs-LC composites, the intensity of the scattering of single GNR is at a maximum when Pk, while it is at a minimum when Pk, in Figs. 4(a) and 4(b) The dark-field microscopy observation confirms the orientation of GNR perpendicular to the director of LC. The angular-dependent scattering spectra in Fig. 4(c) are obtained by averaging signals over the individual GNRs. The intensity of scattering varies from the minimum (when Pk) to the maximum (when Pk). The scattering coefficient with respect to the angle of P and k are shown in Fig. 4(d). The dark-field scattering spectra of individual GNRs verify the negative order parameter of GNRs in LC host.

 figure: Fig. 4

Fig. 4 The dark-field scattering optical microscopy images obtained for unpolarized light passing first through the cell with aligned individual GNRs in LC and then through a linear polarizer (P) either (a) perpendicular or (b) parallel to the rubbing direction k. (c) The measured polarization-sensitive scattering spectra of GNRs for different angles between the linear polarizer and the rubbing direction. (d) The scattering at the longitudinal SPR peak as a function of the angle between (P) and k. The scale bars are 10μm.

Download Full Size | PDF

4. Conclusions

In conclusion, in this work we achieve a bulk alignment of GNRs in nematic lyotropic LC via surface treatment. The aligned GNRs with considerable high concentration (about 6 wt%) show a strong polarization-dependent extinction and scattering. The alignment of colloidal GNRs in a controllable way on a large scale is achieved for the first time and is of interest for fundamental research and technological application. This method, together with applied external magnetic field, is potentially useful for further study of the detailed interaction of nanoparticles with the micelles and can be applied to align various types of nanoparticles in LC hosts. This useful and rather simple method of alignment is compatible with current liquid crystal display technology and potentially useful in electro-optical devices. For example, the GNRs-LC composite can be used as an E-type polarizer, which transmits electric field component parallel to the direction of alignment and absorbs electric field component perpendicular to the direction of alignment [28]. This method of surface-mediated alignment can be extended to fabricate more refined plasmonic devices range from the nanoscale to microscale by more sophisticated methods, such as changing the polymer surface morphology using atomic force microscope [29,30], switching photoresponsive molecular surface monolayer by polarized laser light [31,32], etc. In addition, based on different optical constants and aligned patterns of GNRs derived from the SPR effect, this method can potentially be used to fabricate optical metamaterials, optical cloaking, aligned nano-antenna arrays and plasmonic gratings.

Acknowledgment

This work was supported by the National Nature Science Foundation of China (61178062 and 91233208) and the National High Technology Research and Development Program of China (2012AA030402). We are grateful to Yaoran Sun, Jianwei Tang and Jun Qian for valuable discussions, and Xiaowei Guan and Xinsong Hu for the help in fabrication.

References and links

1. P. Zijlstra, J. W. M. Chon, and M. Gu, “Five-dimensional optical recording mediated by surface plasmons in gold nanorods,” Nature 459(7245), 410–413 (2009). [CrossRef]   [PubMed]  

2. K. J. Stebe, E. Lewandowski, and M. Ghosh, “Materials science. Oriented assembly of metamaterials,” Science 325(5937), 159–160 (2009). [CrossRef]   [PubMed]  

3. K. Hur, Y. Francescato, V. Giannini, S. A. Maier, R. G. Hennig, and U. Wiesner, “Three-dimensionally isotropic negative refractive index materials from block copolymer self-assembled chiral gyroid networks,” Angew. Chem. Int. Ed. Engl. 50(50), 11985–11989 (2011). [CrossRef]   [PubMed]  

4. R. Shenhar, T. B. Norsten, and V. M. Rotello, “Polymer-mediated nanoparticle assembly: structural control and applications,” Adv. Mater. 17(6), 657–669 (2005). [CrossRef]  

5. J. Perez-Juste, B. Rodrigues-Gonzalez, P. Mulvaney, and L. Liz-Marzan, “Optical control and patterning of gold-nanorod–poly (vinyl alcohol) nanocomposite films,” Adv. Funct. Mater. 15(7), 1065–1071 (2005). [CrossRef]  

6. C. Kuemin, L. Nowack, L. Bozano, N. D. Spencer, and H. Wolf, “Oriented assembly of gold nanorods on the single-particle level,” Adv. Funct. Mater. 22(4), 702–708 (2012). [CrossRef]  

7. R. J. Macfarlane, B. Lee, M. R. Jones, N. Harris, G. C. Schatz, and C. A. Mirkin, “Nanoparticle superlattice engineering with DNA,” Science 334(6053), 204–208 (2011). [CrossRef]   [PubMed]  

8. J. Kao, K. Thorkelsson, P. Bai, B. J. Rancatore, and T. Xu, “Toward functional nanocomposites: taking the best of nanoparticles, polymers, and small molecules,” Chem. Soc. Rev. 42(7), 2654–2678 (2013). [CrossRef]   [PubMed]  

9. M. R. Jones, K. D. Osberg, R. J. Macfarlane, M. R. Langille, and C. A. Mirkin, “Templated techniques for the synthesis and assembly of plasmonic nanostructures,” Chem. Rev. 111(6), 3736–3827 (2011). [CrossRef]   [PubMed]  

10. H. K. Bisoyi and S. Kumar, “Liquid-crystal nanoscience: an emerging avenue of soft self-assembly,” Chem. Soc. Rev. 40(1), 306–319 (2010). [CrossRef]   [PubMed]  

11. R. Pratibha, K. Park, I. I. Smalyukh, and W. Park, “Tunable optical metamaterial based on liquid crystal-gold nanosphere composite,” Opt. Express 17(22), 19459–19469 (2009). [CrossRef]   [PubMed]  

12. P. S. Kumar, S. K. Pal, S. Kumar, and V. Lakshminarayanan, “Dispersion of thiol stabilized gold nanoparticles in lyotropic liquid crystalline systems,” Langmuir 23(6), 3445–3449 (2007). [CrossRef]   [PubMed]  

13. K. J. Wu, K. C. Chu, C. Y. Chao, Y. F. Chen, C. W. Lai, C. C. Kang, and P. T. Chou, “CdS nanorods embedded in liquid crystal cells for smart optoelectronic devices,” Nano Lett. 7(7), 1908–1913 (2007). [CrossRef]  

14. I. Dierking, G. Scalia, and P. Morales, “Liquid crystal–carbon nanotube dispersions,” J. Appl. Phys. 97(4), 044309 (2005). [CrossRef]  

15. Q. Liu, Y. Cui, D. Gardner, X. Li, S. He, and I. I. Smalyukh, “Self-alignment of plasmonic gold nanorods in reconfigurable anisotropic fluids for tunable bulk metamaterial applications,” Nano Lett. 10(4), 1347–1353 (2010). [CrossRef]   [PubMed]  

16. Q. Liu, B. Senyuk, J. Tang, T. Lee, J. Qian, S. He, and I. I. Smalyukh, “Plasmonic Complex fluids of nematiclike and helicoidal self-assemblies of gold nanorods with a negative order parameter,” Phys. Rev. Lett. 109(8), 088301 (2012). [CrossRef]   [PubMed]  

17. J. Pérez -Juste, L. M. Liz-Marzan, S. Carnie, D. Y. C. Chan, and P. Mulvaney, “Electric-field-directed growth of gold nanorods in aqueous surfactant solutions,” Adv. Funct. Mater. 14(6), 571–579 (2004). [CrossRef]  

18. G. von Maltzahn, A. Centrone, J.-H. Park, R. Ramanathan, M. J. Sailor, T. A. Hatton, and S. N. Bhatia, “SERS-coded gold nanorods as a multifunctional platform for densely multiplexed near-infrared imaging and photothermal heating,” Adv. Mater. 21(31), 3175–3180 (2009). [CrossRef]   [PubMed]  

19. L. J. Yu and A. Saupe, “Observation of a biaxial nematic phase in potassium laurate-1-decanol-water mixtures,” Phys. Rev. Lett. 45(12), 1000–1003 (1980). [CrossRef]  

20. R. Bartolino, T. Chiaranza, M. Meuti, and R. Compagnoni, “Uniaxial and biaxial lyotropic nematic liquid crystals,” Phys. Rev. A 26(2), 1116–1119 (1982). [CrossRef]  

21. Y. Galerne, A. M. Figueiredo Neto, and L. J. Liebert, “Microscopical structure of the uniaxial and biaxial lyotropic nematics,” Chem. Phys. 87, 1851–1857 (1987).

22. Y. Hendrikx, J. Charvolin, and M. Rawiso, “Uniaxial-biaxial phase transition in lyotropic nematic solutions: Local biaxiality in the uniaxial phase,” Phys. Rev. B Condens. Matter 33(5), 3534–3537 (1986). [CrossRef]   [PubMed]  

23. Q. Liu, C. Beier, J. Evans, T. Lee, S. He, and I. I. Smalyukh, “Self-alignment of dye molecules in micelles and lamellae for three-dimensional imaging of lyotropic liquid crystals,” Langmuir 27(12), 7446–7452 (2011). [CrossRef]   [PubMed]  

24. I. H. Bechtold, M. L. Vega, E. A. Oliveira, and J. J. Bonvent, “Surface treatments for lyotropic liquid crystals alignment,” Mol. Cryst. Liq. Cryst. (Phila. Pa.) 391(1), 95–110 (2003). [CrossRef]  

25. I. H. Bechtold, J. J. Bonvent, and E. A. Oliveira, “Dynamical behavior of a nematic lyotropic liquid crystal in flat confined samples,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 65(1 Pt 1), 011704 (2002). [PubMed]  

26. O. O. Ramdane, P. Auroy, S. Forget, E. Raspaud, P. Martinot-Lagarde, and I. Dozov, “Memory-free conic anchoring of liquid crystals on a solid substrate,” Phys. Rev. Lett. 84(17), 3871–3874 (2000). [CrossRef]   [PubMed]  

27. B. Bahadur, Handbook of Liquid Crystals (Wiley-VCH, Weinheim, 1998), Vol. 2A.

28. P. Yeh and M. Paukshto, “Molecular crystalline thin-film E-polarizer,” Mol. Mater. 14, 1–19 (2001).

29. J. H. Kim, M. Yoneya, and H. Yokoyama, “Tristable nematic liquid-crystal device using micropatterned surface alignment,” Nature 420(6912), 159–162 (2002). [CrossRef]   [PubMed]  

30. A. J. Pidduck, S. D. Haslam, G. P. Bryan-Brown, R. Bannister, and I. D. Kitely, “Control of liquid crystal alignment by polyimide surface modification using atomic force microscopy,” Appl. Phys. Lett. 71(20), 2907–2909 (1997). [CrossRef]  

31. W. M. Gibbons, P. J. Shannon, S. T. Sun, and B. J. Swetlin, “Surface-mediated alignment of nematic liquid crystals with polarized laser light,” Nature 351(6321), 49–50 (1991). [CrossRef]  

32. A. Martinez, H. C. Mireles, and I. I. Smalyukh, “Large-area optoelastic manipulation of colloidal particles in liquid crystals using photoresponsive molecular surface monolayers,” Proc. Natl. Acad. Sci. U.S.A. 108(52), 20891–20896 (2011). [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 (4)

Fig. 1
Fig. 1 (a, b) Polarized optical microscopy images of GNR-doped LC for two different rubbing directions (indicated by k) between crossed polarizers (indicated by (P) and (A), showing high-quality long-range alignment. (c) Polarized optical microscopy images of GNR-doped LC with a phase compensator (530nm λ−plate, slow axis γ) inserted to the POM, showing the alignment of nematic LC is along the rubbing direction. The scale bars are 50μm.
Fig. 2
Fig. 2 (a) Schematic illustration of GNR alignment with a negative order parameter in the nematic LC. Red lines indicate the uniform director field of LC host. (b) The experiment setup used to measure the polarization-dependent extinction spectra of aligned GNR-LC composites. (c,d) The transmission optical microscopy images obtained for unpolarized light passing first through the cell with aligned GNRs and then through a linear polarizer (P) either perpendicular (c) or parallel (d) to the rubbing direction k. The insets of (c) and (d) show the images of the whole sample (in a centimeter scale) in two orthogonal cases. The scale bars are 50μm.
Fig. 3
Fig. 3 (a) The measured polarization-sensitive extinction spectra of GNRs for different angles between the linear polarizer and the rubbing direction. (b) The measured extinction value at the longitudinal SPR peak as a function of the angle between the linear polarizer and the rubbing direction.
Fig. 4
Fig. 4 The dark-field scattering optical microscopy images obtained for unpolarized light passing first through the cell with aligned individual GNRs in LC and then through a linear polarizer (P) either (a) perpendicular or (b) parallel to the rubbing direction k. (c) The measured polarization-sensitive scattering spectra of GNRs for different angles between the linear polarizer and the rubbing direction. (d) The scattering at the longitudinal SPR peak as a function of the angle between (P) and k. The scale bars are 10μm.
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.