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

Nematicons in planar cells subject to the optical Fréedericksz threshold

Open Access Open Access

Abstract

We investigate, both theoretically and experimentally, self-trapping of light beams in nematic liquid crystals arranged so as to exhibit the optical Fréedericksz transition in planar cells. The resulting threshold in the nonlinear reorientational response supports a bistable behavior between diffracting and self-localized beam states, leading to the appearance of a hysteretic loop versus input excitation. Our results confirm the role of nematic liquid crystals in the study of non-perturbative nonlinear photonics.

© 2014 Optical Society of America

1. Introduction

The generation of spatial solitary waves (in the following simply solitons) stemming from the balance of diffractive beam spreading and self-focusing was one of the first topics studied in nonlinear optics [1]. In the simplest medium case, i.e. a Kerr medium with a local response, the so-called ”Townes solitons” are intrinsically unstable in two transverse dimensions and subject to beam filamentation and collapse [2, 3]. Several approaches have been investigated to avoid collapse, such as 1D geometries [4], higher order [5] and saturating nonlinearities [6], parametric interactions [7], nonlocality [8, 9]. Due to the inherent nonlinear nature of self-localized beams, a natural question is whether solitons can be bistable versus beam excitation. In his pioneering theoretical work, Kaplan stated that two stable solitons can coexist for the same input power when the optical nonlinear response exhibits a threshold in the relationship between refractive index and beam intensity [10]. Since then, bistability of bright solitons has been theoretically investigated in cubic-quintic systems [11] and in colloidal media [12, 13].

Thanks to the high nonlinear response typical of nematic liquid crystals (NLCs) [14], self-trapping of light beams in this class of materials has been extensively investigated in the last years. Stable (2+1)D spatial solitons, named nematicons, have been demonstrated in several configurations, including voltage biased and unbiased cells (see Refs. [15, 16] for an exhaustive review) and exploiting the reorientational molecular response [14]. The latter is subject to a threshold when the electric field of the incident beam and the NLC molecules are orthogonal, the so called Optical Fréedericskz Transition (OFT) [17]. In this Paper we demonstrate that in NLCs subject to OFT, i.e. exhibiting a second-order transition when excited by plane wave-like beams [18], a finite size beam (appreciably diffracting as it propagates in the medium) can undergo a first order transition when self-trapping is accounted for, eventually leading to the formation of reorientational solitons. Self-focusing was shown to yield hysteretic effects with an external mirror in previous demonstrations with NLCs [19]. Hereby we demonstrate that, beyond a threshold input power, the free energy of the system has two local minima versus maximum reorientation. Thus, the equilibrium point becomes dependent on the previous evolution history, leading to a hysteresis loop versus excitation. Finally, while Braun et al. used a similar configuration in NLC-filled capillaries for light self-trapping [20], our system benefits from an external voltage bias in order to minimize the optical power for OFT, and our planar geometry permits to avoid the instabilities deriving from spontaneous symmetry breaking.

2. Self-trapping of light in the presence of OFT

The basic principle behind the reorientational nonlinearity of (undoped) NLCs is simple: the (electric component of the) impinging optical field induces dipoles in the elongated organic NLC molecules along a preferential direction, as dictated by their anisotropy; these dipoles in turn tend to align to the field vector or polarization direction. The molecules tend to locally align with the same orientation due to strong intermolecular links; the macroscopic (average) alignment direction pointwise is called molecular director and usually indicated by [14]. Hence, light in ordered NLCs propagates in an uniaxial crystal with the optic axis corresponding to and refractive indices equal to n and n for polarizations normal and parallel to , respectively. The net effect of an (intense) electric field distribution, such as a light beam, on the NLC molecules is to reorient the director according to its intensity profile, thereby changing the local extraordinary refractive index ne [15]. Let us consider a planar cell of thickness Lx = 100μm along x filled with the standard nematic liquid crystal E7, infinitely wide along y and extending from z = 0 to infinity (Figs. 1(a)–1(b)). We take the director to be parallel to z everywhere in the absence of excitations. The director distribution is given pointwise by the angle θ formed with , with θm the maximum θ in each section normal to . A linearly polarized Gaussian (TEM00) beam of wavelength of λ = 1064nm propagates paraxially along z and impinges on the sample; thus, its electric field E can be written as E = A exp[ik0ne(θm)z] (k0 = 2π/λ). To help the optical reorientation, a bias voltage V can be applied to the cell across x, providing a quasi-static electric field ELFV/Lx. In the hypotheses above, the director can rotate in the plane xz and θ unambiguously determines the director as well as the extraordinary refractive index distributions. If spatial walk-off is ignored, the beam evolves according to

2ik0ne(θm)Az+Dx(θm)2Ax2+2Ay2+k02Δne2(θ)A=0,
2θ+ε02K(εa|A|22+ΔεLFELF2)sin(2θ)=0.

 figure: Fig. 1

Fig. 1 (a) Side and (b) top view of NLC planar cell; the blue arrows indicate the director distribution at rest. (c) Free energy F versus θm when K = 12×10−12N and the temperature is 18°C. (d) Soliton width versus input power P corresponding to (c). (e) Sketch of the hysteresis loop in the plane (P, θm). (f) Power threshold for a fixed Gaussian beam of waist 3.5μm versus applied bias V (blue line with crosses, left axis); the dotted lines with squares graph the loop width versus V between win = 5, 11, 35μm (from bottom to top, right axis) and a soliton with an average width of 3.5μm.

Download Full Size | PDF

Equation (1) is the nonlinear Schrödinger equation (NLSE) modified to account for the medium anisotropy and the strong perturbation regime; Dx=ne2(θm)/(n2sin2θm+n2cos2θm) is the diffraction coefficient along x, whereas Δne2=ne2(θ)ne2(θm) is the nonlinear index well associated with the induced director rotation. Eq. (2) is cast in the single elastic constant K approximation [14, 15] and describes the director distribution affected by the external stimuli A and ELF, weighted by the corresponding anisotropies εa=n2n2 and ΔεLF, respectively. Thus, the nonlinear response is highly nonlocal owing to the Poisson-like form of the equation, and saturable owing to the sine term of the torque acting on the molecules [15]. Eqs. (1) and (2) model the propagation of self-confined waves in highly nonlocal NLC, namely nematicons [15].

In standard configurations, the nonlinear response governed by Eq. (2) is thresholdless because the angle θ is finite in the absence of external excitations [15], and Eq. (2) transforms into a Poisson equation for weak nonlinear perturbations. Conversely, when the (electric field) polarization of the input beam and the director are initially perpendicular to one another, θm is non-zero only above the OFT: in this case Eq. (2) cannot be linearized near threshold, with the breakdown of the analytical methods usually employed in nematicon-related models [15, 16]. In order to understand the transition in the presence of self-trapping (given by Eq. (1)), we need to compute the free energy of the system versus the maximum perturbation θm. The simplest approach is to find the equilibrium states studying the soliton case, i.e., with the ansatz A(x, y, z) = ϕ(x, y) exp[ik0(nSne(θm))z], assuming θ invariant across z as well. For the sake of simplicity, we calculate the free energy in the unbiased case V = 0: the application of a bias below the electric Fréedericksz threshold reduces the threshold optical power, without affecting the qualitative trend of the free energy [18].

The overall free energy of the system consisting of NLCs and the electromagnetic field is given by the combination of elastic energy K, given by K2(θθ) in the limit of a single elastic constant, and Lagrangian of the light field Lopt, given by 12ε0n2E2+12ε0εa(n^E)212μ0B2. Note that the light-matter interaction, responsible for torque and reorientation, is contained in the term Lopt. Due to the highly nonlocal character of the NLC reorientational response, ϕ can be taken Gaussian, i.e., ϕ=4Z0P/[πne(θm)w2]exp[(x2+y2)/w2] [9], with P and w the soliton power and waist, respectively; hence, the optical reorientation is θ=0.5γZ0Psin(2θm)Σl=1Vl(y)sin[πl(xLx/2)Lx]/ne(θm), with γ = ε0εa/(4K) and Vl(y) = sinc (πl/2) [F(y) + F(−y)] with F(y)erfc[2y/w+πlw/(22Lx)]eπly/Lx [21]. The overall free energy F=Lx/2Lx/2(K+Lopt)dxdy is [21]

Fακ2θm22+2γZ0PπnSw2Lx2Lx2dxcos(2θ)e2x2+y2w2dy+{ε0Z0nSk02K1w2+ε0Z0nSKγZ0cos(2θm)nS}P,
where κ(w)=1/[Σl=1Vl(0)sin(πl/2)] and α=0.5Σp=0[π2V2p+12/Lx+Lx(V2p+1)2]dy, with V′l (y) = dVl(y)/dy.

Figure 1(c) shows that, below a threshold power Pth, F monotonically increases with θm, thus reorientation does not occur and only the linear regime is stable (in the plot Pth ≈ 18mW). When the power overcomes Pth, a local minimum with θm ≠ 0 appears in F, yielding the formation of a shape-preserving nematicon with width dictated by the power P. Besides the minimum, a local maximum corresponds to an unstable nematicon. Fig. 1(d) compares the nematicon width calculated from Eq. (3) and from numerical simulation of Eqs. (1) and (2). Since for a fixed P every point in Fig. 1(c) corresponds to a different beam width [21], it is straightforward to describe the insurgence of hysteresis. We take a power larger than the threshold Pth and fix the input beam width to win. If the beam is wider than the unstable soliton (i.e., the input state is on the left of the local maximum), the beam evolves towards the linear regime θm = 0. Conversely, if the input beam is narrower than the unstable soliton (i.e., the initial state is placed on the right of the local maximum), the beam reaches the minimum and turns into a stable soliton. Thus, Eq. (3) predicts the occurrence of a hysteretic loop versus power: for a given input waist win a power Pthinc exists such that win corresponds to the local maximum. As power increases from zero to Pthinc(win), reorientation and self-focusing do not take place, and light propagates in the linear regime. When P=Pthinc(win), OFT occurs and the beam starts to self-focus, eventually forming a nematicon with width depending on P according to the existence curve in Fig. 1(d). Now, when decreasing the input power, the beam width differs from win. Since the OFT depends on beam width [22], the director remains reoriented at powers lower than Pthinc(win), as long as P > Pth. The feedback required for memory effects is the nonlocal NLC response, which stores in the director distribution the information on previous optical excitations. Thus, a hysteretic loop for Pth<P<Pthinc(win) is predicted (Fig. 1(e)): the two coexisting stable states of the system correspond to a diffracting beam (lower branch, increasing power) and a nematicon (upper branch, decreasing power). The application of a bias lowers Pth (Fig. 1(f)). Finally, since the larger the input width the larger Pthinc(win) is, the loop width widens as win increases (Fig. 1(f)).

3. Observation of bistability

The results shown in Sec. 2 assume (medium and beam) invariance along z, which is unrealistic in actual situations due to scattering losses, breathing and imperfections. Nonetheless, they provide a qualitative description of beam dynamics in actual experiments. First, the threshold power Pthinc, monotonically increasing with win when z-invariance is invoked, shows a local minimum due to the interplay between the diffractive spreading -decreasing for larger win- and the peak intensity -decreasing with win-. Numerical simulations show that the power at OFT versus win initially diminishes, then has a local minimum when win is between 6 and 8μm and eventually grows indefinitely for further increases in win. Hence, in actual experiments, the hysteresis is maximized when win is as small as possible. Second, when self-focusing occurs, the input beam width does not change, as assumed in Sec. 2. Actually, the beam starts breathing (i.e. its size oscillates periodically) around the equilibrium state shown in Fig. 1(d) [16].

Experimentally, we launched a beam of waist ≈ 2μm at the input. For V = 0V nematicons beyond OFT were subject to temporal instabilities [20]. To circumvent this limitation, we biased the cell with a voltage at 1kHz: such additional electric torque helps molecular reorientation, lowering Pth and avoiding instabilities. At the same time, the loop width reduces as V increases (Fig. 1(f)). We found a satisfactory tradeoff for V = 0.92V and used it in all measurements.

The measured beam evolution in the plane yz is shown in Fig. 2. As predicted, the beam underwent a bistable behavior: when power was ramped up from zero, self-localization did not occur below 16.5mW, the latter corresponding to Pthinc in experiments. Starting from the self-confined state and decreasing power, the system conserved memory of the former state through optical reorientation, and self-trapping was maintained down to 14.5mW, corresponding to Pth.

 figure: Fig. 2

Fig. 2 Observation of the hysteresis loop by imaging beam propagation in the plane yz. (a) The input power is 2mW, corresponding to standard diffraction; the power is increased (b and c) until OFT occurs above 16.5mW; then (d) a stable nematicon is formed. Starting from a stable nematicon, the power is ramped down (b’ and c’): the beam remains self-trapped for P > 14.5mW. Further decreases in power lead to linear diffraction (a) as in first half of the cycle. The beam width w normalized to the apparent input width w0 [22] is plotted versus z for (e) P = 15.5mW and (f) P = 16.5mW; blue and red lines correspond to ramp-up (b–c) and ramp-down (b’–c’) halves of the cycle. Dashed and dotted-dashed lines are the beam widths for P = 2mW and 20mW, respectively. The observed states were stable over time intervals of the order of 30 minutes.

Download Full Size | PDF

The OFT is expected to depend on temperature [14]: Fig. 3 shows hysteresis loops versus temperature. The power threshold decreases as temperature increases, whereas the loop shrinks, nearly disappearing at room temperature. Such trend is confirmed by calculations accounting for the three different elastic constants and the temperature dependence of both refractive indices and elastic constants [22, 23]. Although the loop width is maximum at the lowest temperatures, side effects take place, including longer relaxation times reaching several minutes.

 figure: Fig. 3

Fig. 3 Hysteresis of normalized beam width versus power in z = 930μm, at temperatures of (a) 16°, (b) 18° and (c) 23°C. Panel (b) corresponds to the experiment in Fig. 2. (d) Both the experimental (red stars) and the theoretical (black squares) data show that the loop shrinks at higher temperatures. The theoretical results are found with the z-invariant model [22] and effective widths 3.5 and 11μm for self-trapped and diffracting states, respectively.

Download Full Size | PDF

4. Conclusions

Using a Lagrangian approach for the theory and planar cells for the experiments with near-infrared light, we investigated the role of the optical Fréedericksz transition in the nonlinear propagation of finite-size optical beams. We predicted and observed a bistable behavior accompanied by hysteresis as power was ramped up and down, with stable states corresponding to diffracting and self-trapped beams. The role of power, bias voltage and temperature was addressed in detail and found consistent with simplified models. Our findings underline the role of nematic liquid crystals in the study of highly perturbed regimes in nonlinear optics and further illustrate the wealth of phenomena in nonlocal reorientational soft matter.

Acknowledgments

GA acknowledges the Academy of Finland for his Finnish Distinguished Professor project ( 282858).

References and links

1. R. Y. Chiao, E. Garmire, and C. H. Townes, “Self-trapping of optical beams,” Phys. Rev. Lett. 13, 479–482 (1964). [CrossRef]  

2. R. W. Boyd, S. G. Lukishova, and Y. R. Shen, eds., Self-focusing: Past and Present (Springer, 2009). [CrossRef]  

3. B. Shim, S. E. Schrauth, A. L. Gaeta, M. Klein, and G. Fibich, “Loss of phase of collapsing beams,” Phys. Rev. Lett. 108, 043902 (2012). [CrossRef]   [PubMed]  

4. A. Barthelemy, S. Maneuf, and C. Froehly, “Propagation soliton et auto-confinement de faisceaux laser par non linearite optique de Kerr,” Opt. Commun. 55, 201–206 (1985). [CrossRef]  

5. E. L. Falcão Filho, C. B. de Araújo, G. Boudebs, H. Leblond, and V. Skarka, “Robust two-dimensional spatial solitons in liquid carbon disulfide,” Phys. Rev. Lett. 110, 013901 (2013). [CrossRef]   [PubMed]  

6. J. E. Bjorkholm and A. A. Ashkin, “CW self-focusing and self-trapping of light in sodium vapor,” Phys. Rev. Lett. 32, 129–132 (1974). [CrossRef]  

7. W. E. Torruellas, Z. Wang, D. J. Hagan, E. W. VanStryland, G. I. Stegeman, L. Torner, and C. R. Menyuk, “Observation of two-dimensional spatial solitary waves in a quadratic medium,” Phys. Rev. Lett. 74, 5036–5039 (1995). [CrossRef]   [PubMed]  

8. D. Suter and T. Blasberg, “Stabilization of transverse solitary waves by a nonlocal response of the nonlinear medium,” Phys. Rev. A 48, 4583–4587 (1993). [CrossRef]   [PubMed]  

9. C. Conti, M. Peccianti, and G. Assanto, “Route to nonlocality and observation of accessible solitons,” Phys. Rev. Lett. 91, 073901 (2003). [CrossRef]   [PubMed]  

10. A. E. Kaplan, “Bistable solitons,” Phys. Rev. Lett. 55, 1291–1294 (1985). [CrossRef]   [PubMed]  

11. J. M. Christian, G. S. McDonald, and P. Chamorro-Posada, “Bistable Helmholtz solitons in cubic-quintic materials,” Phys. Rev. A 76, 033833 (2007). [CrossRef]  

12. M. Matuszewski, W. Krolikowski, and Y. S. Kivshar, “Spatial solitons and light-induced instabilities in colloidal media,” Opt. Express 16, 1371–1376 (2008). [CrossRef]   [PubMed]  

13. M. Matuszewski, “Engineering optical soliton bistability in colloidal media,” Phys. Rev. A 81, 013820 (2010). [CrossRef]  

14. I. C. Khoo, “Nonlinear optics of liquid crystalline materials,” Phys. Rep. 471, 221–267 (2009). [CrossRef]  

15. G. Assanto, ed., Nematicons: Spatial Optical Solitons in Nematic Liquid Crystals (Wiley, 2012). [CrossRef]  

16. M. Peccianti and G. Assanto, “Nematicons,” Phys. Rep. 516, 147–208 (2012). [CrossRef]  

17. S. D. Durbin, S. M. Arakelian, and Y. R. Shen, “Optical-field-induced birefringence and Fréedericksz transition in a nematic liquid crystal,” Phys. Rev. Lett. 47, 1411–1414 (1981). [CrossRef]  

18. H. L. Ong, “Optically induced Fréedericksz transition and bistability in a nematic liquid crystal,” Phys. Rev. A 28, 2393–2407 (1983). [CrossRef]  

19. I. C. Khoo, “Optical bistability in nematic films utilizing self-focusing of light,” Appl. Phys. Lett. 41, 909–911 (1982). [CrossRef]  

20. E. Braun, L. P. Faucheux, and A. Libchaber, “Strong self-focusing in nematic liquid crystals,” Phys. Rev. A 48, 611–622 (1993). [CrossRef]   [PubMed]  

21. A. Alberucci, A. Piccardi, N. Kravets, and G. Assanto, “Beam hysteresis via reorientational self-focusing,” Opt. Lett. 39, 5830–5833 (2014). [CrossRef]   [PubMed]  

22. N. Kravets, A. Piccardi, A. Alberucci, O. Buchnev, M. Kaczmarek, and G. Assanto, “Bistability with optical beams propagating in a reorientational medium,” Phys. Rev. Lett. 113, 023901 (2014). [CrossRef]   [PubMed]  

23. A. V. Zakharov and R. Y. Dong, “Giant optical nonlinearity in the mesophase of a nematic liquid crystals (ncl),” Phys. Rev. E 64, 031701 (2001). [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 (3)

Fig. 1
Fig. 1 (a) Side and (b) top view of NLC planar cell; the blue arrows indicate the director distribution at rest. (c) Free energy F versus θm when K = 12×10−12N and the temperature is 18°C. (d) Soliton width versus input power P corresponding to (c). (e) Sketch of the hysteresis loop in the plane (P, θm). (f) Power threshold for a fixed Gaussian beam of waist 3.5μm versus applied bias V (blue line with crosses, left axis); the dotted lines with squares graph the loop width versus V between win = 5, 11, 35μm (from bottom to top, right axis) and a soliton with an average width of 3.5μm.
Fig. 2
Fig. 2 Observation of the hysteresis loop by imaging beam propagation in the plane yz. (a) The input power is 2mW, corresponding to standard diffraction; the power is increased (b and c) until OFT occurs above 16.5mW; then (d) a stable nematicon is formed. Starting from a stable nematicon, the power is ramped down (b’ and c’): the beam remains self-trapped for P > 14.5mW. Further decreases in power lead to linear diffraction (a) as in first half of the cycle. The beam width w normalized to the apparent input width w0 [22] is plotted versus z for (e) P = 15.5mW and (f) P = 16.5mW; blue and red lines correspond to ramp-up (b–c) and ramp-down (b’–c’) halves of the cycle. Dashed and dotted-dashed lines are the beam widths for P = 2mW and 20mW, respectively. The observed states were stable over time intervals of the order of 30 minutes.
Fig. 3
Fig. 3 Hysteresis of normalized beam width versus power in z = 930μm, at temperatures of (a) 16°, (b) 18° and (c) 23°C. Panel (b) corresponds to the experiment in Fig. 2. (d) Both the experimental (red stars) and the theoretical (black squares) data show that the loop shrinks at higher temperatures. The theoretical results are found with the z-invariant model [22] and effective widths 3.5 and 11μm for self-trapped and diffracting states, respectively.

Equations (3)

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

2 i k 0 n e ( θ m ) A z + D x ( θ m ) 2 A x 2 + 2 A y 2 + k 0 2 Δ n e 2 ( θ ) A = 0 ,
2 θ + ε 0 2 K ( ε a | A | 2 2 + Δ ε LF E LF 2 ) sin ( 2 θ ) = 0 .
F α κ 2 θ m 2 2 + 2 γ Z 0 P π n S w 2 L x 2 L x 2 d x cos ( 2 θ ) e 2 x 2 + y 2 w 2 d y + { ε 0 Z 0 n S k 0 2 K 1 w 2 + ε 0 Z 0 n S K γ Z 0 cos ( 2 θ m ) n S } P ,
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.