• No results found

University of Groningen Photoresponsive Self-Assembled Systems Cheng, Jinling

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Photoresponsive Self-Assembled Systems Cheng, Jinling"

Copied!
17
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Photoresponsive Self-Assembled Systems Cheng, Jinling

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Cheng, J. (2019). Photoresponsive Self-Assembled Systems. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Photoinduced Helical Inversion of Cholesteric Liquid Crystals

in Cells with Inhomogeneous Thickness

In this chapter, photoresponsive cholesteric liquid crystals were confined in copper TEM grids, by which a liquid crystal film with inhomogeneous thickness (concave LC cells) could be obtained. Notably, the texture of concave CLC cells can be remotely controlled by UV irradiation, which is highly dependent on the thickness of the LC films.

(3)

5.1. Introduction

Traditionally, Cholesteric liquid crystals (CLCs) can form oily streaks or fingerprint textures under planar or homeotropic anchoring conditions.1,2 In the planar texture of the CLCs, the helices adopt a “standing helix” (SH) orientation (Figure 5.1). The helical axis of the material is perpendicular to the surface of the LC cell, and the selective light reflection (Bragg diffraction regime) is observed with a maximum wavelength λmax = np, where n and p are the average refractivity index and the pitch of cholesteric helix, respectively. The helical pitch p is sensitive to external factors, which is the basis for various practical applications of CLC materials (displays, thermal sensors, smart windows and reflective polarizers, mirrorless lasers, etc.).3,4 The homeotropic texture of the CLCs, where the helical axis is in a ‘lying helix’ (LH) orientation, being parallel to the substrate’s surface, is also known to have fingerprint texture.1 The spatial period and uniformity of these textures can be controlled by varying cell thickness d, pitch p, elastic constants of the LC and anchoring coefficients. Upon optimization of the above parameters and the appropriate alignment of the surface by treatment, highly uniform structures with regular and well-oriented patterns (fingerprints) can be achieved.5

Figure 5.1. Typical alignment of CLC systems a) Planner alignment and b) Vertical alignment

and corresponding textures. Reproduced with permission from Ref 1. Copyright 2003 Wiley-VCH, Weinheim.

In most cases, CLCs are prepared with nematic liquid crystal (LC) materials doped with chiral additives, which induce a helical structure.6,7 This type of CLC has found application in, for example, thermography,8,9 reflective displays,10,11 tunable color filters12,13 and mirrorless lasing.14–16 Meanwhile, dynamic, remote and three-dimensional (3D) control over the helical axis of CLCs is highly demanded, and considered as an important challenge.17 Apart from temperature variation and the use of electric or magnetic fields, the cholesteric helix can also be controlled by light as well.18 The optical control offers several advantages, such as quick and smooth control over the geometries and patterns with spatial-temporal precision.19,20 The initial and photo-induced structures have distinct optical properties,

(4)

therefore these photo-addressable CLCs are very important for optical recording systems,21– 23

all-optical displays,24–26 photo-adjustable light reflectors, etc.12,18

In our research group, molecular motors based on overcrowded alkenes which can undergo unidirectional rotation around the central double bond have been investigated extensively as photoresponsive dopants in LC materials (also see Chapter 3 and 4).24,27–30 For instance, when the CLC phase was deposited into planar anchoring-cells, controllable tuning of color over the entire visible region can be achieved by light.28 In another example, CLC materials were dispersed onto the planar-orientated substrate and left them in air. Under this hybrid anchoring condition, rotary movement of the fingerprint texture was observed when the sample was exposed to UV irradiation.29,31 These systems are based on the flat confinements which provide homogenous geometries for the LC materials. The confinement of LCs imposed by surface boundary conditions plays a key role in most LC devices applications. The confinement effect is due to the boundary conditions, which are enforced by either flat substrates or the presence of a curved boundary. Especially, manipulating the orientational features of LC which are confined to submicrometer space, e. g. spheres,32,33 droplets34,35 or cylinders32,36 consequently appears to be an issue of major fundamental and practical interest for application of these mesogenic materials. However, it is still very challenging to measure the anchoring strength from the interfaces, and it becomes even more difficult when the LC is in curved geometries.

Apart from this, the use of LC materials to fabricate micro-lenses by injecting LC material into convex or concave cells has attracted major attention recently.35,37,38 Unlike the LC cell within flat-substrate devices, the convex or concave cells can provide various thickness of LC films. Jákli and coworkers have demonstrated an assembly-driven micro-lens with size controlled by the grid in which the CLCs films are confined.37 The focusing for the LC lens can be modulated by varying the electric field. Compared with traditional optical technologies, this responsive LC microlens does not require mechanical movements to bring objects into focus. In the reported case, the alignment of substrates and the film thickness has played a vital role.

Although a wide range of motor-doped CLC and alignment coatings has been investigated,24,25,28,29,31,39 a sample of an absolute homeotropic alignment has never been studied for its photo-responsive behavior. Furthermore, our studies mainly focused on large scale and flat LC films.24,25,28,29,31,39 Herein, we applied CLCs onto substrates with homeotropic alignment prepared by coating octadecyltrichlorosilane (OTS) on a glass substrate. In addition, air also promotes homeotropic anchoring and the advantage of using air as a “substrate” is that the homeotropic alignment can be reproduced consistently with air at any level of humidity when comparing with solid substrate.40 In addition, we use a copper TEM grid to confine CLCs into films with small size (400 µm x 400 µm), together with a thickness of about 20 μm and a mesh size of a few hundred micrometers. With the above approach, we can obtain stable, flat and homeotropic orientation of the LC within a fully filled grid (homogenous thickness). At the same time, the preparation of non-flat LC films with inhomogeneous thickness (a concave LC cell) could be achieved as well. The focus of this chapter is to describe the investigation of the photo-induced structural transitions of the CLC cells in both flat and concave surface anchoring confined systems.

(5)

5.2 Results and Discussion

5.2.1 Basic Equations2,41,42

For a flat LC sample, assuming the surface-anchoring energy is the same at every point of surface of the LC film and also assuming the alignment axis is the same on both surfaces. The free energy for the bulk nematic LC phase can be expressed as41, 42:

𝐹 = 𝐾11 2 ( ∇∙ 𝒏) 2 +𝐾22 2 (𝒏 ∙ ∇ × 𝒏) 2+𝐾33 2 (𝒏 × ∇ × 𝒏) 2 5.1 Where, K11 is the splay elastic constant, K22 is the twist elastic constant and K33 is the bent elastic constant. If nematic materials are doped with chiral molecules, according to the elastic continuum theory, the free energy is considered to be mainly associating with the twist elastic deformation, which can be expressed as:

𝐹 = 𝐾22 2 𝑞

2 5.2 Where q is the helical wavevector of CLC material, q is positive for a right-handed and negative for a left-handed twist. We found for the nematic host E7, K22 = 7.4 pN and K33 = 16.5 pN at 25 ºC temperature (Here, we note that the relatively low concentration of chiral dopants in the mixtures used ranging from 0.1 wt% to 1 wt% is not expected to substantially modify the elastic constants of the nematic host).43 At low density of the chiral dopants, the wavevector q of the twist of the free CLC is:

𝑞 =2𝜋𝑝 5.3 where 𝑝 is the initial helical pitch and is determined by the helical twisting power (HTP) of the chiral dopants as

𝑝 =THP∙𝑐∙𝑒𝑒1 5.4 where c is either weight concentration or molar fraction of the chiral dopant, and ee is the enantiomeric excess.

As a result of the photochemical and thermal isomerization of the motor in the LC system, two isomers of the motor are present. As a consequence,44

𝐾22= 𝑥𝑃𝐾𝑃 + 𝑥𝑀𝐾𝑀 5.5 Where KP(M) is the chiral strength contribution either from P isomer or M isomer.

Within the TEM grid, approximating the total energy provided by the grid, substrate and bulk CLC is

𝐸𝑡𝑜𝑡𝑎𝑙 = 𝐹𝐶𝐿𝐶𝑠+ 𝐸𝑠𝑢𝑏𝑠𝑡𝑟𝑎𝑡𝑒+ 𝐸𝑔𝑟𝑖𝑑+ 𝐸𝐿𝐶/𝑎𝑖𝑟 5.6 Accordingly, as the copper TEM grid promotes homeotropic anchoring; there is a periodic director distortion around the cell perimeter. The homeotropic penetration depth h is approximately equal to the half of the cholesteric pitch (h ≈ 𝑝/ 2). This distortion can be estimated by the following formula:45

𝐸𝑔𝑟𝑖𝑑 =𝐾22 2 (

2𝜋 𝑝0)

2 𝐿 𝑑 ℎ 5.7 Where, L=2π/r, r is radius with 𝑟 = 200 𝜇𝑚 in our case. Note that when the CLCs completely wets the wall (d=20 µm, p0=2 μm, K22=8.5 pN), we can get Egrid~1.2∙10-12J/m2. In addition, the

(6)

surface anchoring energy is considered strong in the range of ~ 10-4 J/m2 and weak in the range of ~10-6 J/m2. Since, the ELC/air is 7∙10-6 J/m2.46 The anchoring energy from the wall of grid was found to be negligible for the bulk CLCs system.

The surface anchoring model is especially important to describe the orientation transitions in LC films with an initial homeotropic orientation of the mesogenic molecules at the boundary.47,48 For a flat CLC sample, the lying cholesteric helix can be distorted or even be completely suppressed when CLC is confined in a LC cell with sufficiently strong homeotropic anchoring. There is a certain minimal helical pitch value:

𝑝𝑡ℎ = 2𝑑𝐾22

𝐾33 5.8 If the initial helical pitch p0 is smaller than the threshold pitch pth, the chiral torque is strong

enough with respect to the elastic torque determined by orientational elasticity and anchoring, and twisted cholesteric structure can be observed. When p> pth, the chiral torque

of the CLC material is too weak and the system is not able to twist. Under this condition, the homeotropic structure of NLC cannot be observed.

For concave-shaped CLC samples, the minimums thickness is found in the center of the film, and according to eq. 5.8, the threshold value of pitch pth for the middle of the LC film is

correspondingly less than at other points. As a result, the unwinding of the texture can be first reached in the middle of the film. Ideally, we assume the thickness at the grid edge is around 20 µm, as in the middle of the grid, the stripes of fingerprints can still be observed, as pth~2 µm, the threshold d in the center of the grid is calculated to be 7.5 µm. If HTP decreases, the anchoring energy from the substrate will increase relatively to elastic energy of the CLCs phase. When the helical pitch excessed the pth the texture will fade firstly at the middle of the sample.

5.2.2. Fabrication of Liquid Crystal Cells

The structure of LC lens which is commonly used for the fabrication of micro-lens arrays with an inhomogeneous cell gap is shown in Figure 5.2a.38 The LC lens is constructed by a convex glass lens with a sandwich cell structure. In the present approach, as far as we know for the first time, we constructed the LC cell with concave-shaped simply by applying a copper TEM grid together with the flat substrate as shown in Figure 5.2b

Figure 5.2 Schematic diagram of the microlens array fabrication: a), Conventional fabrication

as reported in the literatures38; b), The design used in the present study and c), Schematic representation of switchable CLC helix based on the photochemical and thermal isomerization of molecular motor (M1)

(7)

The fabrication process for the main substrate requires three major steps with standard substrate deposition and CLC assembly. In a typical procedure, glass substrates were cleaned and coated with OTS by which a perpendicular (homeotropic) orientation of LC molecules can be obtained.46, 49 A 20 μm-thick copper-coated specimen grid was placed onto a chemically functionalized glass slide. The mixture of CLC was made by doping 0.83 wt% of molecular motor M1 in the nematic LC host E7 as shown in Figure 5.2c. A not completely filled (concave) CLCs cell was prepared by filling the pores (400 μm × 400 μm) of the grid until interference rings were observed (the interference rings were monitored by Polarized Optical Microscope (POM) and more details are shown in the Experiment Section). The flat CLCs cells were prepared by dispensing CLCs onto each grid, and the excess of LC was removed latterly (also see in Chapter 4). The filled cells were heated up to 60 °C, and then gradually cooled down to room temperature (1 °C/min). The method results in the formation of a stable film of CLCs within the grid.

5.2.3 Texture of Different CLC Cells under Polarized Optical Microscope (POM)

Stable fingerprint textures were observed under the polarized optical microscope (POM) for all of the CLCs films. When the grid was fully filled with CLC material, uniform light color textures can be observed as shown in Figure 5.3a, indicating a uniform film thickness (around 20 µm). The fingerprint textures in Figure 5.3a are mainly composed by off-centered Moiré-type stripes,37 suggesting that the alignment was not a hybrid alignment but a uniform homeotropic alignment from the bottom to the top layers. The homeotropic anchoring boundary was provided by the interface i.e. air (top) and OTS-coated substrate (bottom) together. Additionally, we found that the size of the stripes is ~1 µm (determined from the period of the fingerprint texture) which is half of the initial pitch of the material.

Figure 5.3 Textures for different CLCs cells on OTS-coated substrate. a), Flat CLC cell using

molecular motor M1 (0.83 wt%) doped in nematic LC host E7; b), Concave-shaped CLCs cell; c), Micheal-Levy chart (reproduced from Olympus-lifescience with permission, copyright Olympus.)40

(8)

Figure 5.3b shows the texture patterns for the concave-shaped CLCs cells. In Figure 5.3b, fingerprint textures were observed with the background birefringence color gradually varying across five different rings. The birefringence color changed from yellow at the grid edge and via green, pale orange, yellow, green, purple to the green at the center of the grid. Due to the limitation of technique in our lab, we used the Micheal-Levy chart40 (Figure 5.3c), to monitor the variation of the thickness of the concave-shaped CLCs cell. It can be seen that birefringence color changed across four orders indicating a 1200 nm change of an optical path difference (Δn x Δd, Δn is the birefringence of the LC sample and Δd is the difference of thickness of the sample). As 0.83 wt% of motors was doped into the nematic E7, and the birefringence Δn referred to the birefringence of E7 as Δn= 0.1, the difference of thickness from the side to the center was calculated to be 12 µm. In other words, the thickness of CLC film at the center of the grid is 8 µm and gradually increases to 20 µm at the edge of grid. Based on these experimental observations, we assumed that a semi-ellipsoidal shaped lens was obtained.

5.2.4 Photoresponsive Behavior

Having established that the CLCs confined in the grid showed stable fingerprint textures, we examined the responsive behavior of these CLCs films when the cells were exposed to both in the flat and concave assemblies.

The non-completely filled (concave) homeotropic LC cell was first tested as demonstrated in Figure 5.4. Surprisingly, upon irradiation for 10 s with 365 nm light, the fingerprint textures started to fade away in the center. It turned to black in accompany with a lengthening of the pitch from initial 2.0 µm (Figure 5.4a) to 3.1 µm (Figure 5.4b) (the pitch was determined from the period of the fingerprint texture) which suggests expanding of the helical pitch. As illustrated in Figure 5.4c-e, the texture gradually disappearing from the middle to the periphery upon continuously UV irradiation until the whole grid appeared to be black. The black image might correspond to a compensated nematic state under the homeotropic alignment which is further confirmed by the conoscopic image (inset in Figure 5.4e).46, 49 The transition time from optical appearance to the dark state that corresponded to homeotropic “pseudo-nematic" LC phase31 was found to be within 10-20 s.

(9)

Figure 5.4 Optical micrographs of the reorganization of curved CLC phases upon UV

irradiation (λ = 365 nm). Below each figure is the homeotropic anchoring condition and its corresponding schematic images: blue rods represent left-hand helix and orange rods represent right-hand helix. The thin film of E7 doped with 0.83% of M-M1 is confined in a TEM grid. From a to b, the pitch increases; from c to e, the helix is unwound to dark state; from f to h, the helix with opposite handiness is rewound; from h to i, the pitch decreases. Scale bar, 100 µm.

Interestingly, if the sample was further exposed to UV irradiation, the texture of the LC film at the grid center was regained as shown in Figure 5.4f. The texture pattern expanded from center to the grid wall until the whole grid was fully covered with the fingerprint textures (Figure 5.4f-g). At the same time the fingerprint patterns with different birefringence colors reappeared from the center to the side of the grid. In addition, during the unwinding and rewinding process of the texture, the expansion or elimination of the texture stripes was observed instead of rotation (rotation of the texture was observed in our previous work 27, 31) indicating that there was no reorientation of the alignment of LC materials during the irradiation. When compared with the patterns of CLCs before (Figure 5.4a) and after irradiation (Figure 5.4i), it can be observed that the concentric stripes have left-handedness

(10)

helix before irradiation and right-handedness afterward. Importantly, at any stage of the irradiation process, the change of the texture can be stopped and reversed by turning off UV irradiation. We have demonstrated here that the optical signature of textures of CLC with homeotropic alignment within a confined environment can be controlled reversibly by light. The distinct unwinding and rewind process during the irradiation can be precisely addressed.

Figure 5.5 Optical micrographs of a glass rod on a CLC film (with homeotropic alignment)

during irradiation with UV light (λ = 365 nm). The glass rod as indicated in the red cycles did not rotate or drift during the process. The dimension of the rod is 28 µm long and 5 µm in diameter. Scale bar, 100 µm.

A micro-glass rod was also dispersed onto the CLCs film as indicated in Figure 5.5. The glass rod was found to neither rotate nor drift during the irradiation (Figure 5.5) suggesting that there is no reorientation in the LC martials. The LC molecules maintain the homeotropic alignment during the irradiation, which further indicates that the black state is caused by the mesogenic molecules with perpendicular orientations.

To further understand the above phenomenon, we look in more detail at the irradiation process. The configuration of CLCs film under homeotropic anchoring is helical, with the helix axis parallel to the substrate and air interfaces. In the absence of UV light, the contribution from P-isomer is negligible (we started with pure M isomer), as the pitch can be expressed as p~c HTPM. Upon irradiation with UV light, the increasing concentration of P-M1

that gradually cancels out the contribution of M-M1 in HTP. The cholesteric pitch increases until the overall optical activity of the fingerprint texture vanishes and the mixture becomes “pseudo-nematic”. We refer to this state as pseudo-nematic because the pseudo-racemic mixture contains chiral isomers having opposite HTP, which means that the resulting effect is equivalent to a racemic mixture where the effects of chirality is canceled out. In addition, under the uniformed homeotropic anchoring conditions, the black state can be observed. As we proceed with UV irradiation, the texture reappears, and at the same time, the chirality is recovered in the optical signature of the CLC material, but with an opposite handedness (Fig. 5.4f-i).

Furthermore, the observed behavior of directional change suggests that the structural transformations in the samples is dependent on the cell thickness (d), which varied from the core to edge of the grid. There are two possible mechanisms, that both the irradiation direction and the thickness of the cell could play a role. First, irradiation form the top might

(11)

induce non-uniform irradiation across the cells. The light beam, passing through the LC film, is attenuated. The thinner part of CLC film at the center of the grid of the curved CLC cells has a relatively higher efficiency for the transformation from M isomer to P isomer, i. e. the “pseudo-nematic” phase can be reached preferentially at the center and therefore the black state can be observed first. The other possible explanation would be, according to the eq. 5.4, that the threshold value of pitch pth increases with the cell thickness. pth has the

minimum value in the middle of the sample when compared with other areas in the film. It is observed that in the middle of the film, the pitch p firstly exceeds the threshold value pth as the helix is completely suppressed and therefore the nematic homeotropic texture (black state) is observed. The fact that the fingerprint textures vanish at first in the middle of the cell indicates the weakening of twisting tension by photoisomerization of M1. Note that at this point, at the molecular level, the “pseudo-racemic” mixture is not reached yet, however, at macroscopic level, the helical organization unwinds and the “pseudo-nematic” phase is resulted. Gradually, the unwinding process spread to the side, as the LCs layer near the grid has the highest pth. Further UV irradiation might lead to the achiral situation at the molecular level as well as the LC phase. As a result, the unwinding process happened at the center firstly, and spread to the side continuously, which can be attributed to the variation of cell thickness.

Further irradiation generates P isomers dominantly, as the helical pitch decreased with the opposite helicity until the photostationary state (PSS) is reached (Figure 5.4f to 5.4i,). Again, photoisomerization of the LC materials at the grid center is more efficient which results in the bright texture first. However, the anchoring boundary is still in favor of maintaining the unwinding state of the LC phase. Based on these observations, we can propose that photoresponsive behavior of the motors has a more pronounced effect than the anchoring effects from the substrate.

To verify our hypothesis, the fully filled homeotropic CLC cell was investigated and the results shown in Figure 5.6. In Figure 5.6a-d, upon irradiation with 365 nm light, a lengthening of the pitch could be observed from the initial 2 µm (a) to 6 µm (d) (the pitch is determined from the period of the fingerprint texture). Upon extended UV irradiation, the period of the surface pattern faded away until the entire image appeared to be black (Figure 5.6f). Further UV irradiation resulted in the reappearance of the texture where the excess P isomer dictates the sign of the overall HTP (Figure 5.6g-i). The synchronous unwinding and rewinding process might be due to the same efficiency (when the LC film has the same thickness, efficiency of photochemical isomerization will be the same) of photoisomerization and consistent anchoring energy across the overall LC material, which further revealed that for the concave CLC cell the directional texture changes is dependent on variation of the thickness of LCs film.

(12)

Figure 5.6 Photoresponsive behaviors of CLCs with uniformed thickness under homeotropic

anchoring condition. The CLC film has uniformed thickness (~20 μm), and the size of the each copper grid is 400 µm x 400 μm.

5.2.5 Thermal Behavior

Previous studies of CLC systems on flat substrates 29,39 have showed that when the irradiation was ceased, the cholesteric textures of M1 doped CLCs films started to rotate in the opposite direction due to the thermal helical inversion (THI) of M1. In the present work, we also investigated the thermal behavior of the CLC cells after the irradiation step.

(13)

Figure 5.7 Optical micrographs of the reorganization of curved CLC film following thermal

isomerization of the molecular motor (P-M1→M-M1) under homeotropic anchoring condition. The CLC film has various thicknesses from the center to the side of the sample. Scale bar, 100 µm.

The thermal isomerization process was examined for the concave-shaped CLC cell with homeotropic alignment after irradiation as indicated in Figure 5.7. As one can see that the bright texture first fades away starting from the center of the sample until the whole grid appeared to be black (Figure 5.7a-d). This phenomenon is attributed to the thermal isomerization of the molecular motors from the P isomer to M isomer. When the P and M isomers are present in equal amount, the homogeneous nematic LC is formed, and under the homeotropic alignment, a black state can be observed. During the study of the thermal behavior, it was observed that the thermal step was much slower than the photoisomerization process. The unwinding process from the bright state (Figure 5.7a) to the entirely dark state (Figure 5.7d) takes 2 min, while the unwinding process induced by photoisomerization of the motor takes 30 s (Figure 5.5a-d). This is in agreement with the kinetics of the thermal process for CLCs as reported in Chapter 3. Note that the efficiency of THI is uniform at every point of CLC film despite of the different thickness of the cell. Therefore, the dark state which originates from the middle of the cell is due to the anchoring energy from the substrate caused by the thickness of LC film. In other words, the initial THI process of the molecular motor induced a lower HTP and longer helical pitch, and the LC materials in the center of the film reached the threshold value of pitch pth first, since the CLC

(14)

Figure 5.8 Optical micrographs of the reorganization of flat CLC film following thermal

isomerization of the molecular motor (P-M1→M-M1) under homeotropic anchoring condition. The CLC film has uniformed thickness (~20 μm), and the size of the each copper grid is 400 µm x 400 μm.

Interestingly, when the sample was kept at the same temperature for a longer period, the reappearance and expansion of the texture from the middle of the cell is not observed. Instead, the process started from the side of the cell and then spread to the center, as shown in Figure 5.7e-g, until the cholesteric texture is fully reestablished. Such phenomenon is possibly induced by the heat flow from the copper grid, where the grid was heated up due to the photothermal behavior.50 As a result the thermal isomerization process was ascended first at the grid wall, where the excess M isomer regenerated the bright texture. The fulfilled LC cell was also investigated as shown in Figure 5.8. As aforementioned, for a flat LC film, the difference from the anchoring boundary for the LC phase is neglected. The cholesteric pitch increased until the overall texture vanished to the black state and the directional texture changes was not observed (Figure 5.8a-e). The reappearance of the textures of the LC film (Figure 5.8e-h) happened simultaneously, but the helical pitch of the LCs at grid wall was much smaller when comparing with the LCs in the middle of the grid, which might be due to the heat flow from the grid. The detailed mechanism of the rewinding process in the concave CLC cell during the thermal isomerization is not yet clear and additional investigation is required.

5.3 Conclusion

In conclusion, we have studied the photo-responsive concave CLC films in confined environment and investigated the influence of the film thickness. When the CLC materials are confined in the TEM grid, a difference of 12 μm in thickness is achieved providing a concave LC lens. A light-induced inversion of helical property accompanied with a smooth and broad-range variation of helical pitch and a sequence of photoinduced structural transitions from “left-handed CLCs to pseudo-nematic LCs to right-handed CLCs” is achieved in both concave and flat CLC films. In addition, after investigating the effect of the surface anchoring, the thickness of the film, the efficiency of isomerization and bulk forces within the CLC, we can conclude that at the center of the curved CLCs where the LC film has minimum thickness, photoisomerization of the motor is more effective than that at the edge

(15)

of LC film, which consequently induced the unwinding and subsequently rewinding process of CLCs in a directed manner (from inside to outside). Furthermore, the thermal isomerization of the molecular motors also involves the unwinding (from inside to outside) and subsequently rewinding process (from outside to inside) in which the two steps have opposite directions. However, the mechanism of these two different processes is still not yet clear. Previous experiments have shown that AFM is a powerful tool to investigate the topography of the surfaces of CLC.29 Therefore, further analysis by different techniques are still needed to address the mechanism of the thermal process of CLC films.

5.4 Experimental Section

Materials. Octadecyltrichlorosilane (OTS) was purchased from Sigma-Aldrich, and used

without further purification. The glass microscope and the capillary glass tube (5 µL) were purchased from Fisher Scientific. Copper TEM grids (thickness is around 20 µm, the pores size is around 400 µm x 400 μm) were purchased from Electron Microscopy Sciences. The details of synthesis procedure and the characterization of molecular motor M1 can be found in Chapter 3.

Treatment of Glass Microscope Slides with OTS.

Glass microscope slides were cleaned according to a procedure detailed in an earlier chapter (Chapter 2). The slides were immersed into a piranha solution (70% (v/v) sulfuric acid and 30% (v/v) hydrogen peroxide) for 1 h at 80 °C (Caution! Piranha solution is highly corrosive and

reactive toward organics). The slides were then exhaustively rinsed with water, ethanol, and

methanol and dried under a nitrogen flow followed by heating at 120 °C for 2 h prior to OTS deposition.

Treatment of glass substrate with octadecyltrichlorosilane (OTS) was performed according to a published procedure.49 A 0.5 mM OTS solution was prepared by adding OTS to heptane that was dried over an aluminum oxide column. The slides were immersed in the 0.5 mM OTS in heptane solution for 30 min at room temperature. Subsequently, they were rinsed with methylene chloride and dried under a nitrogen flow followed by heating. The quality of the OTS slides was evaluated by forming a sandwich of two OTS treated slides; E7 was introduced between the slides and the resulting optical texture was examined using polarized light to confirm homeotropic anchoring.

Preparation of Optical Cells. The copper specimen grids were cleaned sequentially in

methylene chloride, ethanol, and methanol, dried under nitrogen, and then heated at 110 °C for 24 h. The grids were then placed onto the surface of OTs treated glass. One microliter of LC was added into each grid, and the excess LC was removed by contacting a 5 µL capillary tube (Fisher) with the LC droplet on the grid. This procedure led to the formation of a stable film of LC within the grid. The optical cell was heated to ∼60 °C and then cooled down to the room temperature at a speed of 1 °C/min.

Optical Examination of LC Textures. The orientation of LC was examined by using

plane-polarized light in transmission mode on an LV100 POL polarizing microscope (Nikon). The cells were placed on a rotating stage located between the polarizers. Homeotropic alignments for E7 were determined by first observing no transmission of light over a 360° rotation of the stage. Insertion of a condenser below the stage and a Bertrand lens above the stage allowed conoscopic examination of the cell. UV irradiation was performed by a M365F1 LED with 1.0 mW power. The light source was held at an angle of 60° with respect

(16)

to the sample plane, to allow irradiation under the microscope. The distance between the lamp and the sample was approximately 15 cm.

5.5 References

1 I. Dierking, Textures of liquid crystals, Wiley-VCH, Weinheim, Germany, 2003.

2 H. Kelker and R. Hatz, Handbook of liquid crystals, Wiley-VCH: Weinheim, Germany, 1980.

3 B. T. Hogan, E. Kovalska, M. F. Craciun and A. Baldycheva, J. Mater. Chem. C, 2017, 5, 11185–11195.

4 H. K. Bisoyi and S. Kumar, Chem. Soc. Rev., 2011, 40, 306–319.

5 K. Takatoh, M. Sakamoto, R. Hasegawa, M. Koden, N. Itoh and M. Hasegawa,

Alignment technology and applications of liquid crystal devices, CRC Press: London,

2005.

6 R. Eelkema and B. L. Feringa, Org. Biomol. Chem., 2006, 4, 3729–3745.

7 G. Solladié and R. G. Zimmermann, Angew. Chem. Int. Ed., 1984, 23, 348–362.

8 J. T. Crissey, J. M. Bettenhausen and J. L. Fergason, J. Invest. Dermatol., 1965, 45, 329– 333.

9 L. Gao, Y. Zhang, V. Malyarchuk, L. Jia, K. I. Jang, R. Chad Webb, H. Fu, Y. Shi, G. Zhou, L. Shi, D. Shah, X. Huang, B. Xu, C. Yu, Y. Huang and J. A. Rogers, Nat. Commun., 2014, 5, 1–10.

10 J. Kobashi, H. Yoshida and M. Ozaki, Nat. Photonics, 2016, 10, 389.

11 S.-T. Wu and D.-K. Yang, Reflective liquid crystal displays, Wiley New York, 2001, vol. 471496111.

12 N. Y. Ha, Y. Ohtsuka, S. M. Jeong, S. Nishimura, G. Suzaki, Y. Takanishi, K. Ishikawa and H. Takezoe, Nat. Mater., 2008, 7, 43.

13 Z. Zheng, Y. Li, H. K. Bisoyi, L. Wang, T. J. Bunning and Q. Li, Nature, 2016, 531, 352– 356.

14 H. Coles and S. Morris, Nat. Photonics, 2010, 4, 676.

15 G. Cipparrone, A. Mazzulla, A. Pane, R. J. Hernandez and R. Bartolino, Adv. Mater., 2011, 23, 5773–5778.

16 L. Chen, Y. Li, J. Fan, H. K. Bisoyi, D. A. Weitz and Q. Li, Adv. Opt. Mater., 2014, 2, 845– 848.

17 E. Sackmann, J. Am. Chem. Soc., 1971, 93, 7088–7090. 18 H. K. Bisoyi and Q. Li, Chem. Rev., 2016, 116, 15089–15166. 19 M. Russew and S. Hecht, Adv. Mater., 2010, 22, 3348–3360.

20 Ben L. Feringa and W. R. Browne, Molecular Switches, Wiley-VCH,Weinheim, Germany, 2001.

21 L. Wang and Q. Li, Chem. Soc. Rev., 2018, 47, 1044–1097. 22 K. Ichimura, Chem. Rev., 2000, 100, 1847–1874.

23 V. K. S. Hsiao, Y. B. Zheng, B. K. Juluri and T. J. Huang, Adv. Mater., 2008, 20, 3528– 3532.

24 R. Eelkema and B. L. Feringa, Chem. Asian J., 2006, 1, 367–369.

25 R. A. van Delden, N. Koumura, N. Harada and B. L. Feringa, Proc. Natl. Acad. Sci., 2002,

99, 4945–4949.

(17)

& Sons, 2012.

27 J. Vicario, M. Walko, A. Meetsma and B. L. Feringa, J. Am. Chem. Soc., 2006, 128, 5127–5135.

28 R. A. van Delden, N. Koumura, N. Harada and B. L. Feringa, Proc. Natl. Acad. Sci., 2002,

99, 4945–4949.

29 R. Eelkema, M. M. Pollard, J. Vicario, N. Katsonis, B. S. Ramon, C. W. M. Bastiaansen, D. J. Broer and B. L. Feringa, Nature, 2006, 440, 163.

30 J. Chen, E. Lacaze, E. Brasselet, S. R. Harutyunyan, N. Katsonis and B. L. Feringa, J.

Mater. Chem. C, 2014, 2, 8137–8141.

31 R. Eelkema, M. M. Pollard, N. Katsonis, J. Vicario, D. J. Broer and B. L. Feringa, J. Am.

Chem. Soc., 2006, 128, 14397–14407.

32 G. P. Crawford and S. Zumer, Liquid crystals in complex geometries: formed by

polymer and porous networks, CRC Press, 2014.

33 F. M. Van Der Kooij and H. N. W. Lekkerkerker, Nature, 2000, 406, 868–871.

34 I.-H. Lin, D. S. Miller, P. J. Bertics, C. J. Murphy, J. J. De Pablo and N. L. Abbott, Science, 2011, 332, 1297–1300.

35 M. Humar and I. Muševič, Opt. Express, 2010, 18, 26995–27003.

36 H. S. Kitzerow, B. Liu, F. Xu and P. P. Crooker, Phys. Rev. E, 1996, 54, 568–575.

37 P. Popov, L. W. Honaker, M. Mirheydari, E. K. Mann and A. Jákli, Sci. Rep., 2017, 7, 1–9. 38 H.-C. Lin, M.-S. Chen and Y.-H. Lin, Trans. Electr. Electron. Mater., 2011, 12, 234–240. 39 J. Vicario, M. Walko, A. Meetsma and B. L. Feringa, J. Am. Chem. Soc., 2006, 128,

5127–5135.

40 P. Popov, E. K. Mann and A. Jákli, J. Mater. Chem. B, 2017, 5, 5061–5078.

41 H. Yokoyama, Handbook of Liquid Crystal Research, Oxford University Press, New York, 1997.

42 P. J. Collings and J. S. Patel, Handbook of liquid crystal research, Oxford University Press, 1997.

43 J. F. Strömer, E. P. Raynes and C. V Brown, Appl. Phys. Lett., 2006, 88, 51915.

44 A. Bosco, M. G. M. Jongejan, R. Eelkema, N. Katsonis, E. Lacaze, A. Ferrarini and B. L. Feringa, J. Am. Chem. Soc., 2008, 130, 14615–14624.

45 A. E. Rock, Liquid crystal interfaces: experiments, simulations and biosensors, PhD thesis, Kent State University, 2013.

46 J. K. Gupta, M.-V. Meli, S. Teren and N. L. Abbott, Phys. Rev. Lett., 2008, 100, 48301. 47 P. Ribiere, S. Pirkl and P. Oswald, Phys. Rev. A, 1991, 44, 8198.

48 I. I. Smalyukh, B. I. Senyuk, P. Palffy-Muhoray, O. D. Lavrentovich, H. Huang, E. C. Gartland Jr, V. H. Bodnar, T. Kosa and B. Taheri, Phys. Rev. E, 2005, 72, 61707.

49 J. M. Brake and N. L. Abbott, Langmuir, 2002, 18, 6101–6109.

50 G. Baffou, P. Berto, E. Bermúdez Ureña, R. Quidant, S. Monneret, R. J. Polleux and H. Rigneault, ACS Nano, 2013, 7, 6478–6488.

Referenties

GERELATEERDE DOCUMENTEN

It is observed that the chiral motor dopants could induce reorganizations of the LC phase, and the changes in HTP during UV irradiation and thermal relaxation

First, the concentration of chiral motors has a major influence on the order of LC phase at the aqueous-LC interface as higher concentration of the doped molecular motors induced a

We reasoned that this system represents a metastable helical structure and that subsequent thermal helix inversion of the less stable (M,M)-cis-L2 motor unit to the stable

Although compound M3 does not show a liquid crystalline phase (which is possibly attributed to the steric hindrance in the packing of the mesogen) the two

Nature has intrigued and inspired man for millennia with its beauty and fascinating molecular systems allowing complex functions to be performed. In the past few

De natuur heeft de mens al duizenden jaren geïntrigeerd en geïnspireerd met zijn schoonheid en fascinerende moleculaire systemen waardoor complexe functies kunnen

I’d like to thank some of my Chinese friends in Groningen who make me feel less

intrinsic LCs properties, which might be due to the steric hindrance, the distinct isomers are still effective as chiral dopants in nematic LC materials. Overall, the