University of Central Florida STARS

Electronic Theses and Dissertations, 2004-2019

2014

High Performance Three-Dimensional Display Based on Polymer- Stabilized Blue Phase Liquid Crystal

Yifan Liu University of Central Florida

Part of the Electromagnetics and Photonics Commons, and the Optics Commons Find similar works at: https://stars.library.ucf.edu/etd University of Central Florida Libraries http://library.ucf.edu

This Doctoral Dissertation (Open Access) is brought to you for free and open access by STARS. It has been accepted for inclusion in Electronic Theses and Dissertations, 2004-2019 by an authorized administrator of STARS. For more information, please contact [email protected].

STARS Citation Liu, Yifan, "High Performance Three-Dimensional Display Based on Polymer-Stabilized Blue Phase Liquid Crystal" (2014). Electronic Theses and Dissertations, 2004-2019. 4563. https://stars.library.ucf.edu/etd/4563 HIGH PERFORMANCE THREE-DIMENSIONAL DISPLAY BASED ON POLYMER-STABILIZED BLUE PHASE LIQUID CRYSTAL

by

YIFAN LIU B.S. Tsinghua University, China, 2007 M.S. Ohio State University, USA, 2009

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in CREOL, The College of Optics & Photonics at the University of Central Florida Orlando, Florida

Summer Term 2014

Major Professor: Shin-Tson Wu

© 2014 Yifan Liu

ii ABSTRACT

Autostereoscopic 2D/3D (two-dimension/three-dimension) switchable display has been attracting great interest in research and practical applications for several years. Among different autostereoscopic solutions, direction-multiplexed 3D displays based on microlens array or barrier are viewed as the most promising candidates, due to their compatibility with conventional 2D display technologies. These 2D/3D switchable display system designs rely on fast switching display panels and photonics devices, including adaptive focus microlens array and switchable slit array. Polymer-stabilized blue phase liquid crystal (PS-BPLC) material provides a possible solution to meet the aforementioned fast response time requirement.

However, present display and photonic devices based on blue phase liquid crystals suffer from several drawbacks, such as low contrast ratio, relatively large hysteresis and short lifetime.

In this dissertation, we investigate the material properties of PS-BPLC so as to improve the performance of PS-BPLC devices. Then we propose several PS-BPLC devices for the autostereoscopic 2D/3D switchable display system designs.

In the first part we evaluate the optical rotatory power (ORP) of blue phase liquid crystal, which is proven to be the primary reason for causing the low contrast ratio of PS-BPLC display systems. Those material parameters affecting the ORP of PS-BPLC are investigated and an empirical equation is proposed to calculate the polarization rotation angle in a PS-BPLC cell.

Then several optical compensation methods are proposed to compensate the impact of ORP and to improve the contrast ratio of a display system. The pros and cons of each solution are discussed accordingly.

iii In the second part, we propose two adaptive focus microlens array structures and a high efficiency switchable slit array based on the PS-BPLC materials. By optimizing the design parameters, these devices can be applied to the 2D/3D switchable display systems.

In the last section, we focus on another factor that affects the performance and lifetime of

PS-BPLC devices and systems: the UV exposure condition. The impact of UV exposure wavelength, dosage, uniformity, and photo-initiator are investigated. We demonstrate that by optimizing the UV exposure condition, we can reduce the hysteresis of PS-BPLC and improve its long term stability.

iv

Dedicated to my parents.

v ACKNOWLEDGMENTS

My advisor Prof. Wu always tells us: “It is a blessing if you can finish your PhD program in 5 years without changing the research topics.” Indeed, every student who finished his/her PhD program smoothly must have received lots of help and support from many people, and he/she should be grateful to all of them.

First I would like to express my gratitude to my advisor Professor Shin-Tson Wu for his supervision and support to me during the past five years. It is his help and caring that paved my way through the PhD program, and encouraged me to walk through it.

I am also grateful to all my advisory committee members, Professor Pieter Kik, Professor

Patrick Likamwa, Professor Lei Zhai and Professor James Harvey for their insightful comments and helpful suggestions to my research projects.

I want to thank Professor Hongwen Ren from Chonbuk National University (Republic of

Korea) and Dr. Su “Tracy” Xu for their training and help. They are not only my officemates and colleagues, but also my best friends.

I would also like to thank Yifen “Even” Lan from AU Optronics, for his useful discussion and inspiring suggestions to my research in blue phase liquid crystals.

I also appreciate Professor Yan Li’s (Shanghai Jiaotong University, China) support and encouragement to my research while she was in our group, and to my life in Orlando.

In addition, I would like to thank all my colleagues in Display and Photonics Group in

CREOL. It is a memorable experience to work with them, to receive help from them, and to become friends with them.

vi I also want to express my appreciation to the financial support from Industrial Technology

Research Institute (ITRI) in Taiwan, AU Optronics in Taiwan, Air Force Office of Scientific

Research (AFOSR) and Office of Naval Research (ONR).

Finally, this dissertation is dedicated to my family. Even though they are physically away from me, their hearts and minds always accompany me, through my whole life.

vii TABLE OF CONTENTS

LIST OF FIGURES ...... x

LIST OF TABLES ...... xiv

LIST OF NOMENCLATURE ...... xv

CHAPTER 1 INTRODUCTION TO THREE DIMENSIONAL DISPLAY AND

POLYMER-STABILIZED BLUE PHASE LIQUID CRYSTAL ...... 1

1.1 Autostereoscopic 2D/3D Switchable Display...... 1

1.2 Polymer-Stabilized Blue Phase Liquid Crystal...... 8

CHAPTER 2 PS-BPLC MATERIAL FOR DISPLAY APPLICATION: IMPACT OF

OPTICAL ROTATORY POWER ...... 14

2.1 Optical Rotatory Power of BPLC ...... 14

2.2 Compensation Methods of BPLC Optical Rotatory Power ...... 26

CHAPTER 3 PS-BPLC DEVICE DESIGN FOR 2D/3D DISPLAY SYSTEM ...... 37

3.1 High Efficiency BPLC Adaptive Lens Designs...... 37

3.1.1 PS-BPLC Adaptive Lens Design Based on Cascaded BPLC Layers ...... 38

3.1.2 PS-BPLC Adaptive Lens Design Based on Resistive Film ...... 47

3.2 High Transmittance PS-BPLC Slit Design for 2D/3D Display Application ...... 54

3.2.1 2D/3D Switchable Display System Based on IPS PS-BPLC Scanning Slit Array...... 54

3.2.2 High Transmittance IPS PS-BPLC Slit Design ...... 59

CHAPTER 4 OPTIMIZATION OF PS-BPLC DEVICE FABRICATION PROCESS: UV

EXPOSURE CONDITION...... 68

4.1 PS-BPLC Cured at λ=385 nm ...... 69 viii 4.2 PS-BPLC Cured by λ=365 nm...... 71

4.3 Impact of Photoinitiator ...... 75

4.4 Hysteresis-free PS-BPLC...... 78

CHAPTER 5 SUMMARY ...... 80

APPENDIX: LIST OF PUBLICATIONS ...... 82

LIST OF REFERENCES ...... 86

ix LIST OF FIGURES

Figure 1-1. Examples of [5] ...... 2

Figure 1-2. Image reconstruction in electro- ...... 3

Figure 1-3. 2D/3D switchable display based on lenticular adaptive LC lens array [7] ...... 4

Figure 1-4. 3D display based on parallax barrier [13] ...... 6

Figure 1-5. 3D display based on scanning mirror array [15] ...... 6

Figure 1-6. 3D display based on grating array [16] ...... 7

Figure 1-7. Integral image 3D display [18] ...... 8

Figure 1-8. In-plane-switching mode BPLC display device [42] ...... 10

Figure 1-9. Vertical-field-switching mode BPLC display device [39] ...... 11

Figure 1-10. BPLC adaptive lens based on curved electrode [64] ...... 12

Figure 1-11. BPLC adaptive lens based on multiple electrode [65] ...... 13

2 Figure 2-1. (Color online) Temperature dependent ORP (left) and (Δns) (right) of HTG-23 ..... 18

Figure 2-2. (Color online) Bragg reflection wavelength vs. optical rotatory power: red open circles are the measured ORP at =633 nm, green filled circles at =514 nm, and blue triangles at =457 nm. Lines are fitting curves using Equation 4 with only one parameter o=3.12 deg./m.

...... 19

Figure 2-3. (Color online) Light leakage vs. analyzer rotation angle. Here, 0o means the analyzer is crossed with polarizer, and positive rotation angle represents rotation in right-hand direction.

Red open circles are the measured light leakage at =633 nm, green filled circles at =514 nm, and blue triangles at =457 nm. Dashed lines are the overall leakage of white light, including 30% red, 60% green and 10% blue lights...... 21

x Figure 2-4. (Color online) LC director configuration in BP-I phase: Double-twist cylinders ..... 23

Figure 2-5. The multi-domain multi-layer TN cell structure used in the simulation. Upper part: top view, and lower part: cross-section view...... 24

Figure 2-6. (Color online) Polarization state change of the light passing through a multi-domain multi-layer TN cell: black dotted lines show the incident polarization state. Blue solid line shows the simulated outgoing polarization state...... 24

Figure 2-7. (Color online) Simulated isocontrast contour of a BPLC sandwiched between crossed polarizers ...... 28

Figure 2-8. Polarizer and compensation film configuration of the rotated analyzer design: (a) without compensation film; (b) with one biaxial film ...... 29

Figure 2-9. (Color online) Simulated isocontrast contour of a BP LCD shown in Figure 2-8: (a)

Rotated analyzer, without viewing angle compensation; (b) rotated analyzer, with one biaxial film ...... 30

Figure 2-10. (Color online) Simulated isocontrast contour of a BPLC sandwiched between two broadband wide-view circular polarizers ...... 32

Figure 2-11. Compensation configuration based on dispersive +A films ...... 33

Figure 2-12. (Color online) Poincaré Sphere showing polarization state changes when the light traverses through +A films and BPLC layer ...... 34

Figure 2-13. (Color online) Simulated isocontrast contour of BPLC with two dispersive +A films, as shown in Figure 2-11 ...... 36

Figure 3-1. (Color online) Side view of the proposed adaptive BPLC lens ...... 38

Figure 3-2. (Color online) VT curves of BPLC samples: black solid curve for sample I, red dashed lines for sample II and blue dash-and-dot lines for JNC sample ...... 42

xi Figure 3-3. (Color online) Cross-section view of BPLC layers: green convex region is sample I, and yellow concave region is sample II. Black lines are equal potential lines...... 43

Figure 3-4. (Color online) Phase profiles of proposed adaptive lens: (a) using the HTG samples at 30Vrms, 50Vrms, and 100Vrms and (b) using glass to replace BPLC-1, driven at 100Vrms. Black solid curves are for o-wave, blue dashed curves are for e-wave, and red dotted curves are ideal parabolic shapes...... 45

Figure 3-5. (Color online) Voltage dependent focal length of BPLC samples: black solid curve is focal length for o- and e-waves using HTG materials; and blue dashed curves are focal length of o- and e-waves using JNC materials...... 46

Figure 3-6. Side view of the proposed adaptive BPLC lens ...... 48

Figure 3-7. Simulated phase profiles (λ=550nm) across the lens for e-wave (solid curves) and o- wave (dashed curves) at Vo=50Vrms.The upper curve group is for cell gap d=10µm lens, the middle one is for d=15µm, the lower one is for d=25µm...... 50

Figure 3-8. Simulated phase profiles (λ=550nm) across the lens with a 18µm cell gap at 10Vrms,

30Vrms and 50Vrms. Solid curves are for e-wave, dashed curves are for o-wave, and dotted curves are ideal parabolic curves...... 52

Figure 3-9. Simulated voltage dependent focal length of the proposed BPLC lens: blue solid curve is for o- wave, and red dashed curve is for e- wave...... 53

Figure 3-10. Scanning slit based parallax barrier ...... 55

Figure 3-11. Ghost image caused by diffraction effect of IPS electrodes ...... 56

Figure 3-12. Comparison of images with/without IPS BPLC cell placed in front of the screen .. 56

Figure 3-13. Scanning slit array operates in 3D and 2D modes, respectively ...... 57

Figure 3-14. Emission and ambient reflection suppression in an OLED display panel ...... 58

xii Figure 3-15. Simulated equal potential lines of protrusion IPS without (a) and with (b) floating electrodes ...... 61

Figure 3-16. Simulated V-T curves with different electrode configurations. λ=550nm...... 64

Figure 3-17. V-T curves with different cell gaps...... 65

Figure 3-18. V-T curves with different misalignment ...... 66

Figure 4-1. (Color online) Two UV exposure directions: sample 1 is exposed from top substrate side (no electrode) and sample 2 is from bottom side (IPS electrode). LUV λ=385nm ...... 70

Figure 4-2. (Color online) Measured VT curves of samples 1~4: Blue dashed lines are for sample

1, red dotted lines are for sample 2, black solid lines are for sample 3, and magenta dash-and-dot lines are for sample 4...... 71

Figure 4-3. (Color online) Measured transmission spectrum of BPLC precursor in an isotropic state (blue solid line) and BP-I state (red dashed lines) ...... 73

Figure 4-4. (Color online) Polymerization of BPLC cells under different UV exposure conditions

...... 74

Figure 4-5. (Color online) Measured VT curves of three PS-BPLC samples: (a) black solid line is for sample 3 (no photoinitiator); (b) blue dashed lines are for sample 5 (0.2% photoinitiator); (c) red dotted lines are for sample 6 (2.1% photoinitiator). UV lamp: λ=365nm, top exposure.

Probing wavelength λ=633nm ...... 77

Figure 4-6. (Color online) Aging test of a PS-BPLC sample cured under optimized curing conditions. Blue open squares show the hysteresis variation during the test. Red open circles represent the peak voltage variation of the sample during the test...... 79

xiii LIST OF TABLES

Table 2-1. Polarization rotation effect:comparison of IPS and VFS samples. T: Transmittance with analyzer crossed. T’: Transmittance with analyzer rotated by an angle ...... 16

Table 2-2. Simulation parameters ...... 27

Table 3-1. Recipes of the employed BPLC mixtures ...... 40

Table 3-2. Physical properties of the BPLC material we formulated ...... 62

Table 3-3. Protrusion IPS structure parameters ...... 63

xiv LIST OF NOMENCLATURE

2D Two Dimensional

3D Three Dimensional

CR Contrast Ratio

DTC Double twist cylinder

LC Liquid Crystal

LCD Liquid Crystal Display

IPS In Plane Switching

ITO Indium Tin Oxide

PS-BPLC Polymer-stabilized Blue Phase Liquid Crystal

ORP Optical Rotatory Power

TN Twisted Nematic

VFS Vertical Field Switching

xv CHAPTER 1 INTRODUCTION TO THREE DIMENSIONAL DISPLAY AND POLYMER-STABILIZED BLUE PHASE LIQUID CRYSTAL

1.1 Autostereoscopic 2D/3D Switchable Display

3D (Three-dimensional) display has been a hot topic among display industry in the past few years. 3D display feeds the audience with image depth information, which is useful not only for entertainment but also for medical applications and so on. The 3D feeling is generated by feeding two individual images (or two channels of video signals) to two eyes of each audience, and various 3D display approaches have been proposed based on this principle. Those methods requiring the audience to wear glasses or helmet are called stereoscopic 3D displays [1,2], whereas the autostereoscopic techniques do not require any wearing devices [3].

There are three major types of autostereoscopic 3D displays: volumetric display, 3D electro-holography, and direction-multiplexed display. Volumetric display means that the image is filled in a 3D volume [4]. Some examples are sketched in Figure 1-1 [5]. In Figure 1-1(a) and

(b), a single screen is rotated at high speed, scanning through a 3D volume space. So when a 2D image is displayed on this rotating screen, it fills the entire 3D volume. Another example is shown in Figure 1-1(c) where multiple transparent 2D screens are stacked together, so the 2D images displayed on these screens are merged by human eyes into a 3D image.

1

Figure 1-1. Examples of volumetric display [5]

Volumetric 3D system displays the real 3D images for the audience. However, it suffers from several drawbacks, which prevent its commercialization. The first problem is that many objects in the real world are opaque, but the 3D images displayed in the volume are always transparent, so they look not so true. Besides, the volumetric 3D display technology is incompatible with conventional 2D display systems, so the video recording, data processing and signal transmitting systems have to be upgraded to feed the volumetric 3D display systems, which requires a huge investment not only from the customers, but also from hardware manufacturers, TV stations, internet service providers, etc.

The second type of autostereoscopic technology is electro-holography, which displays full color electro-holograms on a 2D screen, and then reconstructs the 3D images using coherent light sources [6]. The mechanism is described by Figure 1-2.

2

Figure 1-2. Image reconstruction in electro-holography

In Figure 1-2, the electro-hologram plate displays the dynamic interfering fringes generated by computers or video recorders. A collimated reconstruction beam illuminates the hologram to reconstruct a 3D virtual image, which is observed by the audience.

Electro-holography could also generate full-color true 3D images. However, in order to achieve a reasonable resolution of the 3D virtual image, the resolution of interference fringes displayed on the hologram plate has to be much higher than the resolution of conventional 2D images. The data processing and the real-time display of these ultra-high-resolution videos are big challenges for current hardware manufacturers.

The third type of autostereoscopic 3D is direction-multiplexed display, in which multiple images are displayed on a 2D screen, either simultaneously or in time sequence. The left eye image and right eye image are separated and guided into left eye and right eye of the audience, respectively. Then the brain of the audience extracts the image distance information from the difference between the left eye image and right eye image. As each individual image entering

3 one eye is still a 2D image, and the 3D feeling is generated by “cheating” the brain, direction- multiplexed displays are sometimes referred to as the “virtual 3D” displays in contrast to the

“real 3D” systems. However, direction-multiplexed display has several attractive features, such as compatibility with conventional 2D display technologies. In principle, each image displayed on the screen is still a 2D image, so the video recording, signal processing and data transmitting hardware are similar to the conventional equipment. The 3D display panel is only an “upgraded” version of a traditional 2D display panel, so the production cost is low, and conventional 2D videos could still be displayed on the same screen very easily. Therefore, direction-multiplexed

3D display is viewed as the most promising candidate for 2D/3D switchable displays.

Several direction-multiplexed 3D display approaches have been proposed, such as [7-12], parallax barrier [13,14], micromirror array [15], grating array [16] and integral image [17-22].

Figure 1-3. 2D/3D switchable display based on lenticular adaptive LC lens array [7]

4 Figure 1-3 is an example of 2D/3D switchable display system based on switchable lenticular microlens array. In this system, nematic liquid crystal is filled into curved lenticular chambers. The refractive indices of the chamber substrates are matched with the ordinary refractive index (no) of the liquid crystal. Therefore, when no voltage is applied, there is a mismatch between the extraordinary refractive index (ne) of the LC and refractive index of the substrate, and each chamber appears as a positive lenticular microlens for the traversing extraordinary light. The light emits from one conventional 2D display pixel under one microlens, and then gets converged towards a predefined direction by the microlens, entering either the left eye or the right eye of an audience. Thus, 3D effect is generated if proper images are displayed simultaneously on the same screen by different groups of 2D pixels. One the other hand, if a 2D image needs to be displayed, a driving voltage is applied on the LC microlens array to reorient the LC molecules. Then the refractive indices of the substrates are matched with no of the LC layer; the focusing power of the LC lens is turned off, and the light emitting from each pixel behind the lens array propagates to all directions, similar to the case in a conventional 2D display panel.

5 Figure 1-4. 3D display based on parallax barrier [13]

Figure 1-4 shows the operation principle of a parallax barrier 3D display system. A parallax barrier layer covers a conventional 2D display panel, so the light emitting from one pixel on the screen is confined by the barrier layer to a specific direction, entering either left eye or right eye of an audience. Therefore, two 2D images are displayed on the same screen at the same time, and then split by the parallax barrier and feed two eyes, respectively. If a 2D video is played, the parallax barrier is turned fully transparent, and the system operates the same as a conventional 2D display panel.

Figure 1-5. 3D display based on scanning mirror array [15]

Figure 1-5 depicts the operation process of a scanning-mirror-array-based 3D display system. Four frames of 2D image are projected onto the micromirror array in time sequence by

6 the projector. The tilting angle of each micromirror is adjusted in accordance with the displayed image. So each image frame is reflected by the micromirror array into an individual eye of the audience.

The 3D display system based on the grating array is similar to the system based on a micromirror array, as shown in Figure 1-6 below. There is a micro-grating array covering the conventional 2D LCD panel. The alignment direction and period of each micro-grating are well- designed, so the light from one pixel on the LCD panel is guided by the corresponding grating towards a predefined direction. Therefore, the images displayed on the same LCD screen by different pixels are separated by the grating array into different directions.

Figure 1-6. 3D display based on grating array [16]

7 Figure 1-7. Integral image 3D display [18]

Figure 1-7 depicts an integral image system. Similar to the 3D display system based on lenticular microlens array, integral image system also has a microlens array, with each microlens covering multiple pixels of a 2D display panel. Each group of multiple pixels displays a small portion of an image, and all the image portions are reconstructed by the microlens array into a

3D virtual image of the object. Comparing to the 3D system based on lenticular microlens array, integral image system provides better image distance feeling and smooth transition between different viewing points, but the limited image depth and low resolution are still open problems.

As a brief summary, multiple autostereoscopic 3D display techniques have been proposed in the past two decades. Among these solutions, the 2D/3D switchable display system based on lenticular microlens array and parallax barrier seems to be the most promising candidate for commercialization, due to its compatibility with conventional 2D display technologies, simple fabrication process and acceptable 3D image quality.

1.2 Polymer-Stabilized Blue Phase Liquid Crystal

Blue phase is a liquid crystal phase in which the liquid crystal molecules are self- assembled in double twist cylinder (DTC) structures [23]. These DTCs are stacked together in body-center-cubic structures (BP-I), simple-cubic structures (BP-II) or amorphous structures

(BP-III). The temperature range of a pure liquid crystal in blue phase is usually 1~2oC, which is too narrow for practical applications. In 2002, Kikuchi et al. demonstrated a polymer stabilization method. This approach expands the temperature range of blue phase to >60oC, including room temperature [24,25]. Currently, polymer-stabilized blue phase liquid crystal (PS-

8 BPLC) is viewed as a promising candidate for next-generation liquid crystal display and photonic devices [26-34].

The operation principle of PS-BPLC is different from conventional nematic liquid crystals. When an electric field is applied on a nematic LC device, the liquid crystal directors are reoriented according to the electric field direction, provided that the employed LC has a positive dielectric anisotropy [35]. But when a blue phase liquid crystal material is driven by an electric field, it switches from optically isotropic state to optically anisotropic. For a BPLC with a positive dielectric anisotropy (Δε>0), the induced birefringence is parallel to the electric field direction, and it follows extended Kerr model [36]:

( ) ( ( ( ) )) ( 1 )

Here Δns is the saturated birefringence of the BPLC, and Es is the saturation field.

PS-BPLC exhibits a faster response time than conventional nematic LCs, which is highly desired for 3D display [37,38]. For display applications, two types of electric fields can be applied to the BPLC device: one is In-Plane-Switching (IPS) mode, and the other is Vertical-

Field-Switching (VFS) mode [39-41]. The device structure of IPS mode is shown in Figure 1-8

[42]. Voltage is applied between pixel (red) and common (blue) stripe electrodes to generate horizontal electric field between these electrodes. The polarizer and analyzer are aligned in 45o and 135o azimuthal angles, respectively, whereas the stripe electrodes are at 90o or 0o. At voltage-off state, the polarization direction of the incident light is not changed by the isotropic

BPLC layer, so the light does not transmit through the crossed polarizers. But in a voltage-on state, the traversing light accumulates phase retardation inside the BPLC layer, changes its polarization state and transmits through the analyzer.

9

Figure 1-8. In-plane-switching mode BPLC display device [42]

IPS BPLC display devices suffer from some drawbacks, such as high driving voltage, noticeable hysteresis and relatively low optical efficiency. In order to overcome these problems, better liquid crystal materials with lower driving voltage have been developed [25,43-45], and different versions of modified IPS mode have been introduced, such as protrusive electrode, corrugated electrode, double-penetration electrode, enhanced protrusive electrode, and so on

[42,46-50].

VFS is another operation mode proposed to solve the aforementioned problems [39]. The structure of a VFS mode device is shown in Figure 1-9. The light is coupled into the BPLC cell at an oblique angle. In the voltage-off state, BPLC is optically isotropic, and does not change the polarization state of the passing light. As a result, the light cannot transmit through the crossed analyzer, which leads to a dark state. When a vertical electric field is applied to the BPLC layer, it induces birefringence in the longitudinal direction. The obliquely incident light accumulates phase retardation, changes its polarization state and propagates through the analyzer. The outgoing light is finally coupled out of the BPLC cell and turned towards normal direction by a coupling film and a turning film, respectively.

10

Figure 1-9. Vertical-field-switching mode BPLC display device [39]

VFS mode makes use of electric field more efficiently than IPS mode, so the driving voltage is much lower, and there is no dead zone. However, due to the propagation of light in oblique angle inside the BPLC cell, the viewing angle of VFS mode display devices is intrinsically limited, and the wide-view compensation films are complicated and expensive, whereas the turning film/coupling film also adds extra cost.

PS-BPLC could also be used in photonic devices, such as adaptive focus lenses. Several types of adaptive lenses have been proposed based on nematic liquid crystal materials. These lenses have tunable optical power, compact size, simple driving scheme and low power consumption [51-59]. But the slow response time and polarization dependency are challenging problems for most adaptive lenses based on nematic LCs [60-62]. PS-BPLC, however, might provide a solution to these problems [63-66]. One example of lenticular BPLC lens array is shown in Figure 1-10. The top electrode has a curved shape, whereas the bottom electrode is planar. Therefore, the spacing between top and bottom electrodes is not uniform. When the voltage is applied between these two electrodes, inhomogeneous vertical electric field is generated across the BPLC cell. The electric field is stronger near the lens edge and weaker in

11 the center. This electric field is primarily in the vertical direction, so the light traversing thro ugh

BPLC in normal direction experiences ordinary refractive index no (E)=niso-Δn(E)/3, which is related to the electric field strength, but regardless of the polarization direction of the incident light. Thus, the refractive index distribution, the optical phase profile and the focal length of the

BPLC lens are determined by the driving voltage. And this BPLC lens is polarization independent. The drawbacks of this BPLC lens design include high driving voltage due to the voltage shielding effect of the polymer layer, and the cost to fabricate curved ITO electrode with high precision, as the electrode curvature determines the phase profile and aberration of the

BPLC lens.

Figure 1-10. BPLC adaptive lens based on curved electrode [64]

12 Figure 1-11. BPLC adaptive lens based on multiple electrode [65]

Figure 1-11 shows another example of BPLC adaptive lens design. There are multiple stripe electrodes on the inner surface of the top glass substrate. The bottom electrode is a planar

ITO electrode coated on the bottom substrate. Each top electrode applies an individual driving voltage, so as to generate gradient vertical electric field distribution across the BPLC layer. The driving voltage applied on each electrode can be optimized to guarantee a high quality parabolic phase delay profile. The dielectric layer is coated on top of stripe electrodes, so that the lateral electric field between top stripe electrodes is shielded, and the vertical electric field distribution becomes smoother, too. However, this BPLC lens design requires complicated driving scheme, and the dielectric layer shields a large portion of vertical electric field, so the driving voltage of this lens is quite high.

According to the discussions above, it is obvious that blue phase liquid crystal has great potential for both display and photonics applications. Therefore, it is worthwhile to consider whether the good features of BPLC could be combined together and applied to 3D display system design. In next chapter, we focus on the optical rotatory power of BPLC material, so as to improve the contrast ratio of BPLC display system. Then in Chapter III, we discuss how to use

BPLC to design new photonic devices and 2D/3D display systems. Afterward, in Chapter IV we optimize the UV exposure process of BPLC devices so as to improve the stability, lower the hysteresis and elongate the lifetime of the BPLC devices.

13 CHAPTER 2 PS-BPLC MATERIAL FOR DISPLAY APPLICATION: IMPACT OF OPTICAL ROTATORY POWER

2.1 Optical Rotatory Power of BPLC

As described in Chapter I, macroscopically, a PS-BPLC is commonly treated as an optically isotropic medium [24,25] without driving voltage. Based on this assumption an excellent dark state (or very high CR) should be obtained when a BPLC cell is sandwiched between two crossed polarizers. However, in the reported BPLC display systems, the demonstrated contrast ratio is as low as 1000:1, which is far below expectation [26,67]. As the brightness of BPLC display in the “on state” is comparable to conventional nematic LCD, it is believed that the low contrast of BPLCD is due to strong light leakage in the “off-state”. Some factors affecting the dark state light leakage have been investigated, such as Bragg reflection and light refraction on the edge of ITO (indium tin oxide) electrodes [67,68]. Undoubtedly, these are important factors, but still another fundamental mechanism should be taken into consideration, which is optical rotatory power (ORP) of BPLC [69-72]. According to Meiboom’s model [73],

BPLC is a three-dimensional structure consisting of double-twist cylinders whose diameter is around 100nm. Inside each cylinder, LC molecules form double-twist structure. When the incident linearly polarized light traverses these cylinders, a small ORP could be accumulated depending on the cylinder’s orientation. As a result, the polarization state of the outgoing light could be rotated by a small angle, which in turn leaks through the crossed polarizer and degrades the CR significantly.

In this section, we characterize the ORP of PS-BPLC samples with different Bragg reflection wavelengths (B). We prove that some double-twist cylinder orientations within a

14 BPLC composite indeed cause the polarization axis of the transmitted beam to rotate and leak through the crossed polarizers, resulting in the degradation of CR. Rotating the analyzer in azimuthal direction to correct the ORP boosts the white light CR of the device by 3-5X. A modified De Vries equation based on a thin twist-nematic (TN) layer is proposed to explain this phenomenon [74].

Our BPLC precursors consist of four ingredients: 1) nematic LC host (HTG135200-100),

2) chiral dopant (R5011); both are from HCCH China; 3) di-functional reactive monomer

RM257 (Merck), and 4) mono-functional reactive monomer C12A (Sigma Aldrich). Detailed preparation process has been reported in Ref. [75]. To obtain different Bragg reflection wavelengths, we varied the weight ratios of LC host and chiral dopant while fixing the monomer weight ratios at 6wt% RM257 and 4wt% C12A. The precursors were cured in the BP-I phase with a UV light (~365 nm and intensity ~8 mW/cm2) for 10 minutes.

The experimental setup for characterizing the optical rotary power of a BPLC cell is described as follows: Three laser wavelengths were used for measurement: R=633nm, G=514nm and B=457nm. The transmission axis of the linear polarizer was fixed, while the analyzer was crossed to the polarizer initially. The BPLC sample was sandwiched in-between, and the light leakage measured. Without a sample, the extinction ratio of the crossed polarizer exceeds 105:1.

We first investigated the BPLC cell with in-plane switching (IPS) electrodes. The electrode width/gap was 8μm/12μm, and cell gap d was 7.4μm. The upper half of Table 2-1 lists the measured transmittance (T) of this sample at the specified RGB wavelengths, when the analyzer was crossed. Surprisingly, T is still quite noticeable. But if we rotate the analyzer in azimuthal direction by a small angle , a much smaller T’ can be achieved for each wavelength,

15 implying the outgoing light still keeps a fairly good linear polarization, but its polarization axis is rotated by .

Table 2-1. Polarization rotation effect:comparison of IPS and VFS samples. T: Transmittance with

analyzer crossed. T’: Transmittance with analyzer rotated by an angle .

λ  /d  (deg.) (nm) (%) (deg./μm) (%)

IPS cell, d=7.4 μm

633 0.02 0.5 0.07 0.004

514 0.10 2.0 0.27 0.020

457 0.60 4.0 0.54 0.090

VFS cell, d=10 μm

633 0.02 0.5 0.05 0.004

514 0.20 2.5 0.25 0.020

457 1.0 5.5 0.55 0.080

This polarization rotation could originate from both BPLC material and light refraction at the edges of electrode; the latter has been reported in Ref. [68]. To rule out the refraction effect and focus on the BPLC material property, we used the vertical-field switching (VFS) cell, which was made of two planar ITO glass substrates without polyimide layer [39,76]. The cell gap was d=10μm. Similar to an IPS cell, we measured the light leakage of the VFS sample. Results are included in the lower half of Table 2-1. We find that crossed analyzer does not lead to the

16 darkest state for either IPS or VFS cell. Rotating a small angle  can reduce the light leakage by

5-10X, depending on the wavelength. The VFS and IPS cells have similar /d and T’ values, indicating that these two cells have a similar optical rotatory property, resulting from the BPLC material property, not from the refraction on ITO electrode edge.

To focus on the material property, in the following experiments we use VFS cells for all the measurements. Considering that outgoing light still keeps a good linear polarization, the polarization rotation angle caused by BPLC is the same as the analyzer rotation angle () from crossed position. Optical rotatory power (ORP)  /d is defined as the polarization rotation angle divided by the cell gap. Unless otherwise mentioned, all measurements were conducted at room temperature (20oC).

In experiment, we first investigate how the temperature affects ORP. We prepared a

BPLC sample, designated as HTG-23, using the above mentioned recipes. Its Bragg reflection

o wavelength is B~410nm and clearing point after UV curing is 78 C. We measured its ORP at

=514nm and 457nm between 30oC and 90oC, and results are plotted in Figure 2-1(a) and (b), respectively. In Figure 2-1, circles represent the measured ORP values. The BPLC sample was placed on a rotary mount whose scale is 2o per division, so the precision of our data is 0.5o in azimuthal angle. Therefore, the data shown in Figure 2-1 appear stepwise. A positive ORP implies that polarization rotation is in the right-hand direction.

17

2 Figure 2-1. (Color online) Temperature dependent ORP (left) and (Δns) (right) of HTG-23

To understand the temperature dependent ORP data in Figure 2-1, we measured the temperature dependent induced birefringence of the PS-BPLC, which in principle can be estimated from [77]:

( ) ( 2 )

Here is the saturated birefringence of Kerr-effect-induced isotropic-to-anisotropic transition in the BPLC [36], Δn0 is the extrapolated birefringence of the LC host at T=0, C is the LC host

’ concentration (wt%) in the composite,  is a material parameter, and Tc is the clearing point of

BPLC after UV curing. Using the measured data (not shown here), we obtained =0.227, and

18 Δn0=0.296, 0.321, and 0.347 for =633nm, 514nm, and 457nm, respectively. Dashed lines in

2 Figure 2-1 represent the temperature dependent (ns) . Semi-empirically, we found that ORP

2 correlates with (ns) reasonably well:

( ) ( 3 )

In total we prepared 15 PS-BPLC samples with different λB’s, and measured their ORP at the specified RGB laser wavelengths. Results are plotted in Figure 2-2. In Figure 2-2, red open circles, green solid circles, and blue triangles represent the measured ORP of BPLC samples at

=633nm, 514nm, and 457nm, respectively, while red solid line, green dashed lines and blue dotted lines are the corresponding fitting curves. The employed fitting equation will be discussed later. From Figure 2-2, when λB increases towards the incident light wavelength λ, ORP increases rapidly. However, when λB gets closer to, or longer than λ, ORP becomes difficult to measure because scattering dominates the transmitted light, and it is difficult to determine  accurately.

Figure 2-2. (Color online) Bragg reflection wavelength vs. optical rotatory power: red open circles are the measured ORP at =633 nm, green filled circles at =514 nm, and blue triangles at =457 nm. Lines are

fitting curves using Equation 4 with only one parameter o=3.12 deg./m.

19 The fast rising of ORP with can also be found in cholesteric liquid crystal as described

2 2 by the De Vries equation [78]. De Vries equation has a term ( /B 1) in the denominator, in which represents the Bragg reflection wavelength of a cholesteric LC. Therefore, it is reasonable to assume the equation of ORP in BPLC also has the same term in the denominator.

Finally, based on the above experimental results we select the following equation to fit the ORP of our BPLC samples:

( ) ( 4 ) ⁄

In Equation 4, is a fitting parameter, is the saturated birefringence of BPLC, which is a function of wavelength and temperature, and is the Bragg reflection wavelength.

We use Equation 4 to fit the measured ORP data, as Figure 2-2 shows. With only one fitting parameter ( 0=3.12deg./μm), good agreement between experiment and Equation 4 is obtained.

Figure 2-2 shows that the ORP of BPLC samples is in the 0.1o/μm to 0.7o/μm range.

Since a typical BPLC cell gap is 5-10μm, the analyzer could deviate by 0.5o to 7o from the perfectly crossed position. To see how this small deviation angle affects the device contrast ratio, we tested the BPLC samples HTG-23 again using the same experimental setup: First, we set the analyzer crossed to the polarizer. Then we rotated the analyzer in azimuthal direction to minimize the light leakage for each wavelength. Result is depicted in Figure 2-3.

20

Figure 2-3. (Color online) Light leakage vs. analyzer rotation angle. Here, 0o means the analyzer is crossed with polarizer, and positive rotation angle represents rotation in right-hand direction. Red open circles are the measured light leakage at =633 nm, green filled circles at =514 nm, and blue triangles at =457 nm. Dashed lines are the overall leakage of white light, including 30% red, 60% green and 10% blue lights.

In Figure 2-3, the horizontal axis shows the analyzer rotation angle (0o means the analyzer is crossed) and vertical axis represents the light leakage percentage (in logarithmic scale). Red open circles, green filled circles and blue triangles are light leakage at =633nm,

514nm and 457nm, respectively. From Figure 2-3, the dark state light leakage can be reduced by

~10X for a given wavelength by adjusting the analyzer’s angle. However, each wavelength has its own optimal analyzer angle. For LCDs, a white backlight is usually comprised of 60% green,

30% red, and 10% blue spectral contents. We take these ratios into account and calculate the overall white light leakage as the black dashed lines shown in Figure 2-3. When the analyzer is crossed, white light leakage is ~0.25%, corresponding to CR~320:1, assuming the bright state transmittance is 80% for a typical IPS cell. If we rotate the analyzer by ~2o, the white light

21 leakage drops to ~0.08%, and the CR is boosted to 1000:1. Moreover, based on Equation 4, if we shift from 410nm to 350nm (this is a common practice for a clear BPLC), the overall leakage would decrease noticeably, and the white light contrast ratio would exceed 3000:1.

Next we try to explain the observed optical rotatory power by a thin TN (twisted nematic) model. It is known that in BP-I phase the LC molecules are arranged in double-twist cylinders, as sketched in Figure 2-4, and these cylinders are stacked to form body-centered cubic (BCC) structure. Let us assume the incident light is in the positive z direction in Figure 2-4, the optical activity of each cylinder is illustrated, too. Due to the random alignment of BCC domains, the incident light could propagate along any direction into the cylinders. But for the simplicity of discussion, we focus on the basic configurations, when the incident wave vector is either parallel or perpendicular to the cylinder’s axis. On the left side of Figure 2-4, the incident light propagates along the cylinder’s axis. Because the LC molecules in the cylinder are aligned equally in all directions, the cylinder appears optically isotropic, which does not affect the polarization state of the incident light. However, for the other two cylinders in Figure 2-4, the light propagation directions are perpendicular to their axis. Therefore, as the light traverses the cylinder it will interact with the 90o twisted LC directors. This twisted alignment affects the outgoing polarization of the light. Generally speaking, the outgoing light will be elliptically polarized, with the long axis of this ellipse rotated by a certain angle θ away from incident polarization direction, as shown in Figure 2-4. This phenomenon is similar to the case when the light passing through a very thin TN cell. Thus, we simulate the optical activity of a TN cell by

Jones matrix and see if we can find analogy between TN and BPLC.

22

Figure 2-4. (Color online) LC director configuration in BP-I phase: Double-twist cylinders

Considering that in a VFS cell there are multiple micro-domains of BPLC, aligning in different directions. Besides, the BCC structure of BPLC has multiple layers of double-twist cylinders stacked inside each domain. Therefore, to imitate the optical activity of BPLC precisely, we use a multi-domain, multi-layer TN LC cell model, as shown in Figure 2-5. We divide a 5μm-thick LC cell into N (N>100) domains. Each aligned in a random azimuthal angle, and divided into multiple layers. Each layer is a 100nm-thick TN cell with 90o twist angle in right hand direction. The LC birefringence is 0.2 and the wavelength is 550nm.

23

Figure 2-5. The multi-domain multi-layer TN cell structure used in the simulation. Upper part: top view, and lower part: cross-section view.

Figure 2-6. (Color online) Polarization state change of the light passing through a multi-domain multi- layer TN cell: black dotted lines show the incident polarization state. Blue solid line shows the simulated outgoing polarization state.

We simulate the polarization of the light passing through each TN domain and then take the average of all domains. Results are depicted in Figure 2-6. Here, horizontal and vertical coordinates represent the amplitudes of light in x and y directions. Black dotted lines show the

24 incident light, linearly polarized in x direction. The outgoing light (the solid blue line) is elliptically polarized, but the eccentricity of this ellipse is close to unity, so it looks like a linear polarization with polarization direction rotated by -4.5o. So the ORP in the thin TN model is defined as the rotation angle of the ellipse’s long axis divided by the cell gap. Based on simulation result, this ORP is described by:

( ) ( 5 ) ⁄

here is a constant, is the LC birefringence, and is the Bragg reflection wavelength of this cell, determined by the refractive index and the thickness of each TN layer. Note that

Equation 5 is quite similar to Equation 4 except for the denominator. Actually, Equation 4 is reduced to Equation 5 when λB<<λ. The physical meaning of this approximation is when the

Bragg reflection wavelength is much shorter than the incident wavelength, Bragg reflection is negligible. Under this approximation, the optical activity of BPLC could be explained by this thin TN model.

To sum up, we have investigated the polarization rotation of polymer-stabilized BPLCs.

The effects of birefringence, Bragg reflection wavelength and incident light wavelength on the

ORP of BPLC are evaluated. A thin TN model is proposed to explain the polarization rotation phenomenon. Simulation result based on this model is found to be an approximation of experimental results when B<<. In next section, based on the investigation of ORP, we will propose several compensation methods to improve the contrast of BPLC display systems.

25 2.2 Compensation Methods of BPLC Optical Rotatory Power

Our experimental results in previous section indicate that the dark state light leakage of a typical multi-domain BPLC cell is reduced by 5~10X if the analyzer is rotated by a small angle to compensate the polarization rotation effect induced by the BPLC layer [79]. This evidence implies that polarization rotation effect, rather than scattering or other mechanisms, plays a major role affecting the dark state light leakage and CR of a BPLC display device. And it is possible to significantly reduce this light leakage and increase the CR by using some optical compensation films designed by the Poincaré Sphere method.

In this section, we propose three compensation film designs to compensate the polarization rotation effect of BPLC in order to achieve high contrast in the normal viewing direction. Besides, the requirement for wide viewing angle is also taken into consideration in our designs.

In the first step, we simulate the contrast of an In-Plane- Switching (IPS) BPLC cell sandwiched between two crossed polarizers and set it as our benchmark for comparison.

Extended Kerr Model (Equation (1)) is used to simulate the bright state transmittance of the

BPLC cell [36]. The saturation field of the BPLC material we employed is 11.1V/μm.

Table 2-2 lists the saturated birefringence (Δns) and optical rotatory power (ORP) of

BPLC, and refractive indices of the polarizers at three primary wavelengths. For the BPLC materials analyzed in Ref. [79], these ORP values correspond to Bragg reflection wavelength

λB=380nm. The thickness of each polarizer is 240μm, and the white light extinction ratio of the polarizers is >20,000:1. The BPLC cell gap is 7.5μm, the IPS electrode width/gap is 5μm/5μm, and the electrodes are aligned at 45o to the absorption axis of the polarizer. The peak

26 transmittance of this BPLC cell in normal direction is ~55%, according to TechWiz simulation.

Then the light leakage in dark state and CR are calculated by MATLAB.

Table 2-2. Simulation parameters

BPLC material properties

Wavelength Optical rotatory power Saturated birefringence:

o (nm) ( /μm) Δns

450 0.31 0.20

550 0.10 0.19

650 0.04 0.17

Polarizer properties

Wavelength Refractive index: Refractive index:

(nm) ne no

450 1.5+i*0.0014 1.5+i*4.1*10-5

550 1.5+i*0.0021 1.5+i*3.1*10-5

650 1.5+i*0.0027 1.5+i*2.9*10-5

Figure 2-7 shows the simulated isocontrast contours with a white light: 60% green light

(λ=550nm), 30% red light (λ=650nm), and 10% blue light (λ=450nm). The CR at normal direction is ~1500:1, which is close to the experimental results reported in Ref. [26,67].

27

Figure 2-7. (Color online) Simulated isocontrast contour of a BPLC sandwiched between crossed polarizers

The easiest way to compensate the polarization rotation effect of BPLC is to rotate the analyzer. When the absorption axis of the analyzer is rotated to match the polarization direction of the outgoing light, the lowest light leakage will be achieved. It should be noted that polarization rotation angle is wavelength dependent. Therefore, rotating analyzer could only minimize the light leakage at one wavelength, rather than the entire visible spectrum. For display applications, we choose to minimize the light leakage at =550nm, in order to obtain the highest white light contrast.

28

Figure 2-8. Polarizer and compensation film configuration of the rotated analyzer design: (a) without compensation film; (b) with one biaxial film

Figure 2-8(a) shows a new BPLCD configuration without any compensation film, in which the analyzer is rotated by a small angle θ away from the crossed state. Arrows represent the absorption axes of polarizer and analyzer, respectively. To achieve minimal light leakage at

=550nm, as mentioned above, θ is calculated to be 0.75o. In Figure 2-8(b), one biaxial film is added to improve the viewing angle [42,80]. The thickness of biaxial film is set as 27.5μm, and the refractive indices are: nx=1.51, ny=1.50 and nz=1.505. Here nx axis is indicated by an arrow within the film plane, and nz axis is in the normal direction of the biaxial film. To maximize the viewing angle, the nx axis of the biaxial film should be perpendicular to the absorption axis of the analyzer.

29

Figure 2-9. (Color online) Simulated isocontrast contour of a BP LCD shown in Figure 2-8: (a) Rotated analyzer, without viewing angle compensation; (b) rotated analyzer, with one biaxial film

Figure 2-9(a) and Figure 2-9(b) plot the isocontrast contour of the new BP LCD configuration without and with biaxial film, respectively. In Figure 2-9, the CR in normal direction is about 3700, which is more than twice of the normal contrast in Figure 2-7. Besides,

30 in Figure 2-9(b), the CR=100:1 envelop covers all viewing directions within 85o polar angle. So the configuration in Figure 2-8(b) is a good design for wide view display applications.

An interesting discovery from our simulation is that green light contributes 60% luminance in the bright state, but only causes 5% light leakage in the dark state. Therefore, it is difficult to further improve the CR by reducing green light leakage. In order to achieve a higher

CR, red and blue lights have to be considered comprehensively.

Our second approach is to use Hong’s broadband and wide-view circular polarizers (CPs)

[81]. The CP converts a linearly polarized incident light into circular. Therefore, the polarization rotation effect inside BPLC will not affect the extinction ratio of the CP. Besides, it is designed for broadband applications, so the light leakage at 450nm, 550nm and 650nm are suppressed simultaneously. The third important feature is that this CP has a wide viewing angle, and no additional viewing angle compensation is needed.

Figure 2-10 depicts the simulated isocontrast contour of a BPLC cell sandwiched between these broadband circular polarizers. The CR at normal direction is as high as 12,000:1.

Besides, in the entire viewing zone the CR is over 100:1, satisfying the wide viewing angle requirement.

31

Figure 2-10. (Color online) Simulated isocontrast contour of a BPLC sandwiched between two broadband wide-view circular polarizers

The configuration based on Hong’s broadband wide view circular polarizer demonstrates high contrast ratio and wide viewing angle. However, this design requires four layers of biaxial films, which not only increases the cost but also introduces extra manufacturing error. A good compensation using fewer compensation films is always desired. The A films have been widely used in the viewing angle compensation of nematic LCDs [82]. Here we demonstrate that dispersive positive A film is also a possible solution for the polarization rotation compensation of

BPLCDs, as Figure 2-11 shows.

32

Figure 2-11. Compensation configuration based on dispersive +A films

The BPLC layer is sandwiched between two identical dispersive +A films. The principal axes of these two +A films deviate a small angle from the absorption axis of the polarizer, one in clockwise direction and the other in counter-clockwise direction (±θ in Figure 2-11). The angle θ and the dispersion of the +A film are optimized to compensate the polarization rotation effect of

BPLC, which will be explained in detail below. In addition to the dispersive +A films, one biaxial film is added to widen the viewing angle.

33

Figure 2-12. (Color online) Poincaré Sphere showing polarization state changes when the light traverses through +A films and BPLC layer

The function of dispersive +A film is described by the Poincaré Sphere in Figure 2-12.

The zoomed-in curves on the surface of the sphere are shown in the inset. Curves 1-3 indicate the polarization state change when the linearly polarized light propagates through the first +A film, the BPLC layer and the second +A film in normal direction, respectively.

The linearly polarized incident light is converted into an elliptically polarized light, after passing through the first +A film (Curve 1). Then the polarization rotation process inside BPLC layer is represented by Curve 2, which is along a latitudinal line on the sphere surface. Finally, the polarization state change through the second +A film corresponds to Curve 3. The outgoing

34 light from the second +A film becomes linearly polarized again, and its polarization direction is the same as the incident light. In another word, the polarization rotation effect of BPLC is compensated by the +A films. In the Poincaré Sphere, the lengths of Curves 1 & 3 relate to the birefringence of +A film. On the other hand, the length of Curve 2 is determined by the polarization rotation angle of BPLC, which is wavelength dispersive. Therefore, by selecting a proper dispersion of +A films, it is possible to achieve perfect compensation in the entire visible spectrum.

Based on the mechanism described above, the parameters of dispersive +A films are optimized for the same BPLC material described before. The +A film thickness is 10μm. The small angles between principal axes of +A films and polarizer absorption axis are ±θ=±1.5o. The birefringence of the +A films is 0.008 for λ=450nm, 0.005 for λ=550nm and 0.004 for λ=650 nm.

The parameters of the biaxial film are the same as mentioned ahead. The compensation result is shown in Figure 2-13. A CR=13500:1 is achieved in normal direction. Within 85o polar angle, the CR is over 100:1. The high CR on axis is due to the specifically designed dispersion of the

+A films, which optimizes the compensation result for all three wavelengths simultaneously.

35

Figure 2-13. (Color online) Simulated isocontrast contour of BPLC with two dispersive +A films, as shown in Figure 2-11

These three compensation film designs each has its pros and cons. The first design is simple, but the compensated performance is not very good. The third design shows the best performance, but it is quite complicated. Based on these possible solutions, the contrast of BPLC display would be comparable with commercial nematic LCDs. And this will makes BPLC a more competitive choice for the 2D/3D switchable display system design.

36 CHAPTER 3 PS-BPLC DEVICE DESIGN FOR 2D/3D DISPLAY SYSTEM

3.1 High Efficiency BPLC Adaptive Lens Designs

Liquid crystal (LC) adaptive lenses with tunable focal length have found useful applications in auto-focusing [83,84], 2D/3D switchable displays [7,85-88], and tunable photonic devices [89-95]. The majority of such adaptive lenses employ nematic LCs because of their high birefringence, simple alignment, and low operation voltage. However, to achieve a short focal length a relatively thick LC layer (d30m) [54,58,86,96] and high birefringence LC material are needed. For a 30m LC layer, its response time is in the order of 1s [97]. A high birefringence (n>0.4) LC helps to reduce cell gap and obtain faster response time, but its ultraviolet (UV) stability is a concern [98]. Moreover, most nematic LC lenses are polarization dependent. To overcome this problem, a straightforward method is to stack two devices in orthogonal directions [99].

PS-BPLC material, on the other hand, provides another solution to those problems aforementioned. The adaptive lens designs based on BPLC have been introduced in Chapter I, with their pros and cons being discussed. Then in this section, we would like to propose two new

BPLC adaptive lens designs, so as to meet the requirement of 2D/3D switchable display system design.

37 3.1.1 PS-BPLC Adaptive Lens Design Based on Cascaded BPLC Layers

In this section, we propose a new microlens structure using two cascaded polymer- stabilized BPLC layers. The Kerr constant of the first BPLC layer is much larger than that of the second one. Due to the curved interface of the two BPLC layers, optical power is generated.

Such a microlens has several attractive features: simple structure, good phase profile, polarization independence under normal incident light, and fast response time.

Figure 3-1. (Color online) Side view of the proposed adaptive BPLC lens

Figure 3-1 depicts the side view of the proposed cylindrical BPLC lens. Sandwiched between top and bottom planar ITO glass substrates are a convex BPLC-1 layer and a concave

BPLC-2 layer. BPLC-1 has a similar dielectric constant and refractive index as BPLC-2, but a much smaller Kerr constant. This can be obtained by using the same LC host, but different weight ratios of photo-curable monomers and chiral dopant. The following paragraphs will explain in detail the mechanism of well-controlled BPLC recipe affecting the Kerr constant of

BPLC. Both layers of BPLC precursors are cured under UV light for sufficient time to ensure that polymerization process has been finished. The radius of the lens aperture is R, the central

38 thickness of BPLC-2 is d1. The edge thickness is d2, which is also the overall cell gap of the LC layer. In order to reduce voltage shielding effect, d1 should be made as small as possible.

When no voltage is applied, there is no optical power because of matched index between the BPLC layers. Once a voltage is applied between top and bottom ITO electrodes, uniform vertical electric fields are generated across the lens due to matched dielectric constant. In BPLC-

1, since the Kerr constant is small, the refractive index change is negligible and the normally incident light sees nave=(2no+ne)/3. In BPLC-2, a uniform refractive index change occurs across the lens, following extended Kerr effect [36], as shown by Equation 1 in Chapter I.

In extended Kerr model, the Kerr constant K is related to ∆ns and Es as:

⁄ ( 6 ) ( )

The normally incident light experiences a refractive index no=nave ∆nind(E)/3 in BPLC-2, which decreases with increasing voltage. The phase profile at any radius is calculated based on the optical path lengths in BPLC-1 and BPLC-2. Due to the concave shape and decreased refractive index of BPLC-2 layer, a phase profile like a positive lens is formed.

The dielectric constant matching plays a critical role in this lens design. The first advantage gained from this feature is that voltage shielding effect is minimized: In this lens design, the voltage drop through the curved BP-BP interface makes contribution to the lens power, whereas the voltage drop in other layers is a unfavorable shielding effect, such as in the central part of BPLC-2, whose thickness is d2. Benefitted from the small thickness of d2 and dielectric constant matching of BPLC1-BPLC2, only a small portion of voltage is shielded, in contrast to the strong shielding effect on the polymer layer reported in Ref. [64]. Another advantage of this design is polarization independence. Because the electrodes are planar and the

39 dielectric constants are matched, uniform vertical electric fields are generated with very little horizontal components. By suppressing the horizontal components, the device is polarization independent [64]. This feature will be quantitatively analyzed in the following sections.

We prepared two BPLC mixtures using the same LC host, HTG135200-100 (from HCCH,

China) and high twisting power chiral dopant R5011 (from HCCH). The recipes for sample I and sample II are listed in Table 3-1, where RM257 (Merck) and TMPTA (Sigma Aldrich) are photocurable monomers. The host LC has following physical properties: Δn =0.205 at λ =642nm,

Δε =85 at 1kHz and 21°C. Different amount of R5011 was used to control the pitch length (or

Kerr constant) of the two BPLC mixtures.

Table 3-1. Recipes of the employed BPLC mixtures

HTG R5011 RM257 TMPTA

Sample 78.0% 10.0% 7.2% 4.8% I

Sample 87.0% 5.0% 4.8% 3.2% II

From Table 3-1, the weight ratio of photocurable monomers are different in the two samples. For sample I, it is 12 wt% in total, but for sample II, it is 8 wt%. This is because polymer network also plays an important role in the Kerr constant of BPLC materials. High concentration of polymer network provides rigid binding to the LC molecules in the double twist cylinder structures, and makes them more difficult to reorient. Macroscopically, this means a smaller Kerr constant. On the other hand, different monomer ratios lead to a slightly mismatched dielectric constant. According to our measurement the dielectric constant of sample I is ave1= 25,

40 and for sample II, it is ave2= 34. Dielectric mismatching is unfavorable because it generates horizontal electric field, which in turn degrades the polarization independence property of the lens. But following simulation will show that such a small mismatch is still acceptable.

In order to measure the Kerr constant of these two samples, we injected the precursors into two vertical field switching (VFS) cells with a 5m cell gap. The transition temperature of the precursors between chiral nematic and blue phase during the heating process is 32.7°C for sample I and 68.7°C for sample II. They were cured during cooling process at 33°C and 70°C, respectively, with an UV light (8mW/cm2) for 10min. After the BP mixtures were polymerized, we measured the voltage-dependent transmittance (VT) curves of the two VFS cells using the same method as reported in Ref. [39]. The incident angle is 70o and wavelength is 633nm.

Results are shown in Figure 3-2.

The black solid curve represents the VT curve of sample I, and dashed red lines represent sample II. We can see that the on-state voltage of sample II is ~22Vrms, whereas the peak voltage of sample I is much higher than 40Vrms. This indicates sample II has a much larger Kerr constant than sample I. As discussed before, there are two reasons for this outcome. First is because Kerr constant is proportional to the square of pitch length [100], the larger concentration of chiral dopant in sample I results in a shorter pitch length and smaller Kerr constant compared to sample

II. The second reason is the higher polymer concentration in Sample I, resulting in a more rigid polymer network and lower Kerr constant. After fitting with extended Kerr model, we obtain the saturation birefringence Δns1~0.16 and saturation electric field Es1=34.15V/μm for sample I and

Δns2~0.18, Es2=11.1V/μm for sample II. From Equation 6, we estimate that the Kerr constant of sample I is about one order smaller than that of sample II.

41

Figure 3-2. (Color online) VT curves of BPLC samples: black solid curve for sample I, red dashed lines

for sample II and blue dash-and-dot lines for JNC sample

Also included in Figure 3-2(blue dash-and-dot curve) is the VT curve of a JNC JC-

BP01M BPLC reported in Ref. [27] (ave= 70, Δnsat= 0.142 and Es= 4.15V/μm). By using this large Kerr constant BPLC composite, the operating voltage is greatly reduced. In order to demonstrate the potential of cascaded BPLC lens design, this JNC material is also used in the simulation (next section). We assume that the original JC-BP01M serves as the BPLC-2 of

Figure 3-1, and BPLC-1 uses another hypothetical BPLC material with the same dielectric constant and Δnsat as BPLC-2, but Es= 40V/μm.

With the physical properties of BPLC samples in hand, we carried out simulation first assuming BPLC-1 is HTG sample I, and BPLC-2 is HTG sample II. First, we used the average dielectric constants of BPLC-1 and BPLC-2 to obtain electric potential distribution using

TechWiz. In the simulation, we assume d1= 1m, and d2= 12m. The radius of the microlens is

40m, the total BPLC cell gap is 13m and the shape of BPLC-1 is sinusoidal. Figure 3-3 shows

42 the cross-section view of the cylindrical BPLC lens described above by Figure 3-1 and equal- potential lines simulated by TechWiz.

Figure 3-3. (Color online) Cross-section view of BPLC layers: green convex region is sample I, and yellow concave region is sample II. Black lines are equal potential lines.

Figure 3-3 shows an entire lens with 13m cell gap and 80m width. It is seen that equal- potential lines are distributed uniformly through the entire cell gap, and primarily in horizontal direction, implying that vertical electric field dominates. Actually, the electric field has strongest horizontal component at point “P” near the edge of the lens, as shown in Figure 3-3. And even in this point, the electric field vector deviates from vertical direction by only 1.6o. This result proves that the small mismatch of dielectric constants in Sample I (ave1= 25) and Sample II

(ave2= 34) does not affect the electric field distribution too much.

43 Then we used the extended Kerr model to calculate the induced birefringence in the

BPLC layers and the phase profile across the lens. Figure 3-4(a) shows the simulated phase profiles using HTG samples at =633nm. The driving voltages are 30Vrms, 50Vrms, and 100Vrms.

The black solid (o-wave) and blue dashed (e-wave) curves overlap very well, indicating that the lens is polarization independent. This is due to the close match of dielectric constant and uniform vertical electric fields generated along the lens. Moreover, both curves match closely to an ideal parabolic shape as depicted by the red dotted curves, which means aberration would be suppressed. For comparison, in Figure 3-4(b) we show the phase profile of a similar BPLC lens, but the BPLC-1 material is replaced by glass (ε=5). The severe mismatch of dielectric constant causes much more voltage shielding on the glass layer. Therefore, the phase delay generated by the BPLC deviates from ideal parabolic profile significantly. Besides, the o-wave and e-wave profiles separate noticeably, implying that the horizontal field is much stronger than the design shown in Figure 3-4(a). As a result, the lens is polarization dependent.

44

Figure 3-4. (Color online) Phase profiles of proposed adaptive lens: (a) using the HTG samples at 30Vrms,

50Vrms, and 100Vrms and (b) using glass to replace BPLC-1, driven at 100Vrms. Black solid curves are for o-wave, blue dashed curves are for e-wave, and red dotted curves are ideal parabolic shapes.

The focal length at different voltages was calculated using:

( 7 ) ( )

45 where (E) is the phase difference between the lens center and edge. As shown in Figure 3-5, the black solid curve is focal length for o- and e-waves using HTG materials. The two polarizations always have the same focal length, due to the polarization independence of BPLC lens. To show the better performance of cascaded BPLC lens design with improved material properties, the focal length performance is also simulated based on JNC materials mentioned in the previous section. The blue dashed curves are focal length of o- and e-waves using JCBP01M and the hypothetical BPLC material. In comparison with HTG materials, JNC materials give a shorter focal length and lower voltage.

Figure 3-5. (Color online) Voltage dependent focal length of BPLC samples: black solid curve is focal length for o- and e-waves using HTG materials; and blue dashed curves are focal length of o- and e-waves using JNC materials.

Although the simulated results of the proposed BPLC lens are interesting, the device fabrication is challenging. A possible solution is to use two-step curing process. In the first step,

46 the top substrate is assembled with a parabolic-shape glass mold to form a LC cell. The BPLC-1 precursor is injected into this cell, and UV cured at BP-I phase. After curing, the BPLC-1 is stabilized and its temperature range broadened significantly. The glass mold is then peeled off, and the planar bottom substrate is assembled. So a chamber is formed between cured BPLC-1 and bottom substrate. In the second step, BPLC-2 precursor is injected into this chamber, and cured at BP-I phase. The curing temperature of BPLC-1 and BPLC-2 may be different. But with the stabilization of polymer network, BPLC-1 layer will stay in blue phase under the curing temperature of BPLC-2.

The most challenging step in this fabrication process is to peel off the glass mode without destabilizing the BPLC-1 layer. A possible way is to freeze the sample in liquid nitrogen before peeling off the glass mold. According to our experience, the frozen polymer network is more resistant to the mechanical impact incurred in the peeling off process. The liquid crystal will be crystalized under low temperature, but it would resume blue phase after unfreezing.

3.1.2 PS-BPLC Adaptive Lens Design Based on Resistive Film

In this section, we propose another new adaptive microlens using a polymer-stabilized

BPLC. The BPLC is sandwiched between a planar PEDOT-PSS resistive electrode and a planar

ITO electrode. Parabolic phase profile is obtained in low field region due to the linear Kerr effect in the BPLC layer with only one data addressing. Our design has a larger optical power compared to those shown in [64,65] at the same voltage because there is no shielding effect from any dielectric layer. Simulation results show that the device is polarization independent and it has parabolic-like phase profile in a large tuning range.

47 Figure 3-6 shows the side view of our proposed BPLC lens. On the inner surface of top substrate, there is a center ITO electrode at the center of the lens, and a ring ITO electrode on the edge. The aperture of the lens is further coated with a thin transparent PEDOT: PSS high- resistive film. On the inner surface of the bottom substrate, a planar ITO electrode is coated. The radius of aperture is R, and the cell gap of BPLC layer is d.

Figure 3-6. Side view of the proposed adaptive BPLC lens

In our design, we ground the center ITO electrode and bottom ITO electrode, but vary the voltage on the edge electrode Vo. When Vo=0, there is no electric field and the BPLC remains optically isotropic. For the normally incident light, both ordinary (o) and extraordinary (e) waves experience niso at any radius. Thus, there is no optical power.

When Vo>0, the PEDOT: PSS electrode generates a linearly varying potential from center to edge as:

( ) ( 8 )

48 where r is the radius at the point of interest. As a result, vertical electric fields with gradient intensity are generated across the lens and so is induced birefringence ∆nind. For a normally incident light, both o- and e-waves see an ordinary refractive index no=niso-∆nind/3 [36], which is also radius dependent. Near the center of the lens, the electric fields are weaker due to smaller voltage difference between top and bottom electrodes, thus the induced birefringence ∆nind is smaller and no(r) is larger. On the other hand, near the edge of the lens, the induced birefringence is larger and no(r) is smaller. Therefore, a phase profile with positive lens is formed.

At low field region, the induced birefringence follows Kerr effect as:

( 9 ) where λ is the wavelength and K is the Kerr constant. At radius r, considering the linear potential distribution in Equation 8, the induced birefringence could be further expressed as:

( ) ( ) ( ) ( 10 )

Equation 10 is a parabolic function with respect to radius r. Consequently, the phase

Ψ(r)=2πdno(r)/λ=2πd[(niso-∆nind(r)/3)]/λ would also have a parabolic shape. In such a simple structure we could obtain parabolic phase profile easily.

However, as the electric field further increases, the induced birefringence would gradually saturate according to the extended Kerr effect [36]. Therefore, the phase profile would deviate from the ideal parabolic shape in the high field region.

In order to validate the device concept, we carried out simulations using finite-difference frequency-domain method. We assume the PEDOT: PSS layer has a linear gradient potential from center to edge. We first calculated the electric field distribution in the LC layer, then the induced birefringence based on extended Kerr model and finally the phase for the o- and e-

49 waves. The BPLC we used is Chisso (now JNC) JC-BP01M [39]. It has Es~5.4V/µm and

∆ns~0.14 at λ=633nm and room temperature. For practical applications, we are more interested in λ=550nm. Thus, we extrapolated the ∆ns to 550nm based on the wavelength dispersion model

[101] and found ∆ns~0.149. Since Es does not depend on the wavelength, it remains the same.

Figure 3-7. Simulated phase profiles (λ=550nm) across the lens for e-wave (solid curves) and o-wave

(dashed curves) at Vo=50Vrms.The upper curve group is for cell gap d=10µm lens, the middle one is for d=15µm, the lower one is for d=25µm.

Figure 3-7 shows the simulated phase profiles of the proposed lens structure with different cell gaps at 50Vrms operating voltage and λ=550nm. The widths of center and edge electrodes are both 2µm, and the aperture radius is 160µm. The black solid curve and dashed curve represent the phase profiles of e- and o- waves, respectively, in a structure with d=10µm;

50 the blue solid curve and dashed curve are for d=15µm; and the red solid curve and dashed curve are for d=25µm.

First of all, at the same applied voltage 50Vrms, the 10µm structure has more than 1π phase difference from center to edge, while the 25µm structure has only~0.6π. Although the

25µm LC layer would potentially have much larger phase difference at higher voltage, at this specific voltage, its phase is smaller due to the much weaker electric fields. Meanwhile, inside the 10µm structure, electric fields are so strong that the induced birefringence deviates from the

Kerr relation (Equation 9). As a result, the phase profile also deviates from the ideal parabolic shape. In the 25µm lens cell, the deviation is much smaller so that phase profile is nearly parabolic. Obviously, a larger cell-gap structure promises a better phase profile at the same voltage.

Secondly, the 10µm structure has a better overlap between o- and e- waves. The smaller the cell gap is, the stronger the vertical components of electric fields are. Relatively, the horizontal components are smaller. As discussed in Ref. [64], horizontal electric field components are the causes of polarization dependency, since o- wave would see a decreased refractive index, while e- wave sees an increased one. With a relatively smaller horizontal component, smaller cell gap is more advantageous for polarization independency. Similarly, a larger aperture would result in a better overlap of the o- and e- waves.

So we have chosen an optimized structure with d=18µm and R=160µm, to have a large phase difference, good phase profile shape and polarization independency at the same time. As shown in Figure 3-8, at 50Vrms a phase change ~0.8π is achieved. The o-wave (red solid curve) and e-wave (red dashed curve) overlap with each other very well, indicating the optimized structure is indeed polarization independent. Moreover, they fit with the ideal parabolic shape

51 (red dot curve) well. Thus, a good image quality could be obtained. Obviously, the lower the operating voltage, the more parabolic like shape would be formed, as shown by the 30Vrms and

10Vrms curves, because the induced birefringence follows Kerr effect better in the low field region.

Figure 3-8. Simulated phase profiles (λ=550nm) across the lens with a 18µm cell gap at 10Vrms, 30Vrms

and 50Vrms. Solid curves are for e-wave, dashed curves are for o-wave, and dotted curves are ideal parabolic curves.

Figure 3-9 shows the simulated voltage dependent focal length for the abovementioned optimized structure. The focal length is calculated using Equation 7 mentioned in the previous section [64,65]. The blue solid curve is the focal length for o- wave, and the red dashed curve is for e- wave. As shown in Figure 3-9, the two waves have the same focal length, which further

52 proves the lens is polarization independent. As voltage increases, the focal length decreases. At

60Vrms, the simulated focal length is 44mm.

Figure 3-9. Simulated voltage dependent focal length of the proposed BPLC lens: blue solid curve is for o- wave, and red dashed curve is for e- wave.

A possible extension of present lens structure is to have the edge electrode and bottom electrode grounded, while varying the voltage on the center electrode. Then a negative lens is formed, but the shape of the lens will not be as good as the positive lens.

Both BPLC adaptive lens designs demonstrate fast response time and polarization independency. But high driving voltage is always a significant problem for these lens designs.

Besides, with the simply electrode designs we use, it is difficult to finely tune the phase profile of the BPLC adaptive lens, so as to improve the image quality of 3D display system. Therefore, it is worthy to investigate another possible 3D display mode, which is the parallax barrier display

53 mode. In next section, we will discuss the possibility of applying IPS BPLC device in the 3D display system design based on scanning parallax barrier.

3.2 High Transmittance PS-BPLC Slit Design for 2D/3D Display Application

In this section, a 2D/3D switchable display system design is proposed based on parallax barrier structure. Scanning slit made of IPS-BPLC cell is applied to improve the resolution of the system in 3D mode.

3.2.1 2D/3D Switchable Display System Based on IPS PS-BPLC Scanning Slit Array

The parallax barrier based 3D display system is a promising candidate for autostereoscopic 3D display modes. Its basic mechanism has been discussed in Chapter I .

Comparing to other autostereoscopic 3D display systems, such as lenticular lens based design or integral-imaging, parallax barrier based design generates images with much lower crosstalk between left-eye and right-eye images [3]. This advantage guarantees better image quality and less discomfort for the audience. However, the parallax barrier also suffers from several drawbacks, such as low resolution and low optical efficiency. The commercialization of this technology would be difficult if these problems are not solved satisfactorily.

As Figure 1-4 shows, the 3D effect of parallax barrier is based on the fact that half of the pixels are blocked by the barrier, and only the other half pixels are observed by each eye of the audience. Therefore, in 3D mode the effective 3D resolution of this display panel is reduced by one half. Scanning barrier could be a solution to this problem, as shown in Figure 3-10.

54

Figure 3-10. Scanning slit based parallax barrier

In Figure 3-10, the video is divided into odd and even frames. When these frames are displayed on the screen alternatively, the slit array in front of the screen is shifted correspondingly. Therefore, the 3D pixels appear to be dislocated by half of a pixel pitch when an odd frame is followed by an even frame, and the dynamic resolution of the 3D video is doubled, i.e., equals to the 2D screen resolution [102].

The scanning slit design requires switching capability of the slit barrier layer, which also induces another advantage-2D/3D compatibility. If the entire slit barrier layer is turned transmissive, all the 2D pixels on the screen will be visible to both left and right eyes, and the display system will become a conventional 2D display.

Due to the even and odd frames, scanning slit design requires doubled frame rate on both slit barrier layer and the display panel. Thus, we propose to use an IPS-BPLC cell as the slit barrier device, as BPLC is the very choice to meet the requirement of fast switching speed.

However, there are several concerns about IPS-BPLC serving as the slit barrier layer.

One of them is the diffraction issue. In order to reduce the BPLC driving voltage, the IPS electrode width and gap are ~5m. The periodic IPS electrode structure will diffract the incoming light, which in turn will introduce ghost images. One example of the diffraction effect is shown in Figure 3-11, when an IPS 10/10 BPLC cell is placed close to the audience’s eye.

55 Several ghost images appear when the viewer looks at the computer screen through the IPS

BPLC cell.

Figure 3-11. Ghost image caused by diffraction effect of IPS electrodes

Diffraction effect is one of the major reasons preventing the application of IPS-BPLC cell in active shutter glasses. However, Figure 3-12 demonstrates that diffraction effect will not bother the audience if the IPS cell is placed close to the screen, as required in the slit barrier.

Figure 3-12. Comparison of images with/without IPS BPLC cell placed in front of the screen

Figure 3-12(a) depicts a desktop computer image without any BPLC cell placed in front of the screen, and the inset shows the pixels on the screen. Figure 3-12(b) shows the same

56 desktop image on the screen with an IPS 10/10 BPLC cell sandwiched between the screen and another analyzer. The inset also shows the pixels viewed through the BPLC cell when the cell is turned on with a proper driving voltage. The pixels in Figure 3-12(b) look similar to those shown in Figure 3-12(a), and no ghost image appears overlapping with the original image. This is due to the fact that the distance between the IPS cell and the observer is quite long. Therefore, the higher-order diffracted lights will deviate far away from the 0th-order transmitted light, and will not enter the observer’s eyes. This proves that the IPS-BPLC cell can be used as the scanning slit array in a parallax barrier 3D display system and solve the resolution loss problem.

The proposed 2D/3D switchable display panel based on scanning parallax barrier is depicted in Figure 3-13.

Figure 3-13. Scanning slit array operates in 3D and 2D modes, respectively

The bottom layer in Figure 3-13 is the analyzer, which is the outmost layer of a conventional 2D LCD panel. In 3D mode, if there is no voltage applied to the IPS-BPLC cell the outgoing light will be completely blocked by the second analyzer on the top. If the IPS-BPLC

57 cell is turned on, light will transmit through the second analyzer. Thus, the IPS-BPLC cell serves as a switchable shutter. And by turning on/off different regions of the IPS cell in sequence, we can generate a scanning slit array and recover the 3D image resolution. In 2D mode, on the other hand, the entire IPS-BPLC cell is turned on, so the light traverses the entire cell and the second analyzer, so the 2D image is visible for all audiences.

This scanning slit design based on IPS-BPLC can be applied to LCD or OLED panel.

OLED is emissive so it does not require analyzer to generate different emitting light gray scales.

But it requires a circular polarizer to suppress the ambient light reflection from the OLED structure in order to improve ambient contrast ratio. The operation principle of the circular polarizer is depicted in Figure 3-14.

Figure 3-14. Emission and ambient reflection suppression in an OLED display panel

As shown in Figure 3-14, the linear polarizer and the quarter wave plate (QWP) together form a circular polarizer. Half of the emitting light from OLED panel transmits through this

58 circular polarizer and reaches the audience. On the other hand, incident ambient light is reflected by the mirror and then completely absorbed by the circular polarizer. Thus, ambient light reflection is suppressed and the contrast ratio of OLED display panel under strong ambient light is improved.

Therefore, both LCD and OLED display panels employ a linear polarizer as the outmost layer in their system designs. So our IPS-BPLC cell and the second analyzer can be laminated on top of this polarizer, and serve as the scanning slit, as depicted in Figure 3-13, no matter whether the 2D display panel is LCD or OLED.

In summary, based on the IPS-BPLC scanning slit we designed a 2D/3D switchable display system. Comparing to conventional 3D display based on fixed parallax barrier, our design demonstrates high resolution in 3D mode. It is capable to operate in combination with conventional 2D LCD/OLED display panel, and the IPS scanning slit design is also compatible with current LCD manufacturing process, so the fabrication cost could be well controlled.

3.2.2 High Transmittance IPS PS-BPLC Slit Design

In previous section, it is proved that IPS-BPLC slit array could serve as the scanning parallax barrier in 3D display system. However, the optical efficiency is still another problem to address. As shown in Figure 3-13, the transmittance of the IPS-BPLC slit array is directly related to the overall optical efficiency of the display system in both 2D and 3D modes. However, d ue to the dead zone on top of each IPS electrode, the transmittance of conventional IPS-BPLC slit is limited, which in turn reduces the optical efficiency of the entire 3D system. In this section, we propose to use floating electrodes on the top substrate in conjunction with bottom protrusion IPS

59 electrodes to improve the peak transmittance to over 90%. The cell gap and floating electrode dimension are optimized to maximize the peak transmittance.

In an IPS-based BPLCD, including double penetrating fringing field mode and enhanced protrusion mode [42,48], the lateral electric field between electrodes determines the field- induced birefringence, and so as the transmittance. However, on top of each electrode, the electric field is primarily in vertical direction. Therefore, the incident light does not experience enough phase retardation, and will be blocked by the crossed analyzer. This low-transmittance region is usually referred to as “dead zone”. To suppress dead zones, double sided IPS electrode design has been proposed, in which electrodes are fabricated not only on bottom substrate but also on top substrate [103]. Top electrodes generate lateral electric field to remove the dead zones of bottom electrodes, and vice versa. Therefore, the dead zone areas of bottom electrodes and top electrodes are complemented. A major tradeoff of this design is that it requires two TFTs

(thin film transistors), which not only reduces the pixel’s aperture ratio but also increases the complexity in pixel registration.

Although the “two-TFT design” is complicated, the idea of reducing dead zones by fabricating electrodes on top substrate is still valuable. The challenge is how to let these top electrodes generate tunable lateral electric field without a second TFT.

To overcome this problem, here we propose a floating electrode approach. These floating electrodes are not connected to any fixed voltage signal; its potential is solely determined by the surrounding electric field. As a result, each floating electrode serves as an equal-potential conductor, which affects the nearby electric field distribution. Thus, when the bottom electrodes are driven by a voltage, the potential of top electrode will be changed accordingly, even although

60 it is not directly driven by a TFT. These lateral fields will help to reduce the dead zone areas and improve the overall transmittance of the BPLC cell.

Figure 3-15 depicts the device structure, equal-potential lines, and corresponding transmittance of a protrusion IPS without (a) and with (b) floating electrodes. To simulate the device performance, we used TechWiz simulation program. In Figure 3-15(a) and Figure

3-15(b), red and blue shapes represent protrusion electrodes on the bottom substrate. In Figure

3-15(b), the floating electrodes are drawn as black lines on the top substrate. As Figure 3-15(a) shows, dead zone appears on top of each protrusion electrode.

Figure 3-15. Simulated equal potential lines of protrusion IPS without (a) and with (b) floating electrodes

As Figure 3-15(b) shows, the floating electrodes affect the electric field distribution. In our simulation, all protrusion electrodes are equally separated, each protrusion is symmetric in shape, and floating electrode is placed in equal distance from two neighboring protrusions. In this case, the potential of floating electrode is the median of potentials on two protrusions. Thus, in the local regions between protrusions and floating electrodes, the electric field is strengthened, which increases the induced birefringence of BPLC. Therefore, even though the effective cell

61 gap on top of protrusion is smaller, the transmittance is still improved. This is why we observe smaller dead zone and higher overall transmittance in Figure 3-15(b).

In Figure 3-15, the protrusion electrodes have reverse-sine side wall, which is also a result of optimization. Yoon, et al has investigated how the side wall shape of protrusion electrode affects the peak transmittance of a BPLC cell [50]. As described in Ref. [50], steep sidewall, such as the top of reverse-sine protrusion electrode, bears a higher transmittance than trapezoid or elliptical protrusion, because it generates a stronger lateral electric field as well as weak vertical electric field between electrodes. Therefore, we also use this protrusion design in conjunction with floating electrode.

A tradeoff of floating electrode is the increased operating voltage. It is known that the electric field near an ideal conductor surface tends to be normal to the surface. This phenomenon is also shown in Figure 3-15(b): the equal-potential lines below floating electrodes are bent towards horizontal direction, which means the vertical component of electric field dominates near floating electrode surface. It reduces the phase retardation in protrusion gap area, and decreases the overall transmittance. So floating electrode will increase the driving voltage, as discussed in the following section.

To simulate the induced birefringence of the polymer-stabilized BPLC composite, we use extended Kerr model, and Table 3-2 lists the properties of the BPLC we prepared in our labs.

Table 3-2. Physical properties of the BPLC material we formulated

λ 650nm 550nm 450nm

Δnsat 0.141 0.154 0.169

Es 4.15V/μm

62

The protrusion electrode dimension and cell gap are also critical parameters. Considering the practical manufacturability, we have chosen following parameters for our simulation, as listed in Table 3-3.

Table 3-3. Protrusion IPS structure parameters

Cell gap 4μm

Protrusion electrode width 3μm

Protrusion electrode height 3μm

Protrusion sidewall shape Reversed sine

We first investigate how the gap (G) between protrusion electrodes and the width (W) of floating electrode affect the transmittance. Figure 3-16 shows the simulated voltage-dependent transmittance (VT) curves at λ=550nm. In Figure 3-16, all the black lines represent the VT curves with G=2.4μm, blue lines for G=3.6μm, and red lines for G=4.8μm. Solid lines are protrusion IPS without floating electrode. Dotted-and-dashed lines are for W=3.6μm, and finally dotted lines for W=4.8μm.

From Figure 3-16, we find that when G=3.6μm and W=4.8μm, the peak transmittance reaches 90%, which is ~15% higher than the case without floating electrode. But the peak voltage increases from ~19V to ~23V.

If G is reduced to 2.4μm, a lower driving voltage is achieved. In this case, W=3.6μm is an optimal choice, which provides ~85% peak transmittance. On the other hand, if G is increased to 4.8μm, the peak transmittance of protrusion IPS electrode without floating is already quite

63 high (~80%), so floating electrode does not help too much. With W=4.8μm, the peak transmittance rises to ~87%, but the driving voltage also increases to 27V.

Figure 3-16. Simulated V-T curves with different electrode configurations. λ=550nm

Another important parameter is cell gap. Figure 3-17 shows the V-T curves vs. cell gap.

Here protrusion electrode gap is fixed as G =3.6μm, and W=4.8μm.

64

Figure 3-17. V-T curves with different cell gaps

Figure 3-17 shows that 4μm is the optimal cell gap (black line). This is due to the fact that the transmittance is improved by strengthening electric field near the protrusion top. So when the cell gap increases from 4μm to 5μm (blue line), the separation between floating and protrusion is enlarged. The voltage between floating electrode and protrusion does not change, so the electric field and the induced birefringence near protrusion top are smaller in the case of thicker cell gap. And as a result, the peak transmittance is lowered. If, on the other hand, the cell gap is reduced to 3.1μm (red line), it is difficult for the incident light to accumulate π phase retardation in such a thin cell. So the peak transmittance is lower, and driving voltage is higher for 3.1μm cell gap.

It is easy to notice that floating electrode has to be fabricated on the top substrate, and misalignment between top and bottom substrates will affect the function of floating electrode. So

65 in this part, we simulate the V-T curves of three configurations, and show them in Figure 3-18.

The first (black line) has ideal alignment, the second one (blue line) has 1μm lateral misalignment between two substrates, and the red line shows the case with 2μm misalignment.

The electrode dimensions of these three configurations are the same as in previous section, and cell gaps are all 4μm. The simulated wavelength is 550nm.

Figure 3-18. V-T curves with different misalignment

From Figure 3-18 we see that when there is 1μm and 2μm misalignment, the peak transmittance falls down in almost the same scale, to ~84%, which is still higher than the peak transmittance of protrusion IPS without floating electrode. The peak voltages of misaligned cases, on the other hand, are almost unchanged.

As a brief summary, we demonstrate that floating electrode is helpful to improve the transmittance of IPS-BPLC display devices, without introducing a second TFT into each pixel.

66 Therefore, we can applied this concept into the scanning slit design of parallax barrier 3D display system, and improve the optical efficiency of the entire system significantly.

67 CHAPTER 4 OPTIMIZATION OF PS-BPLC DEVICE FABRICATION PROCESS: UV EXPOSURE CONDITION

As described in previous chapters, most existing problems of PS-BPLC devices for

2D/3D switchable display applications have already been addressed satisfactorily, such as driving voltage, optical efficiency, contrast ratio and so on [31,46,104]. However, hysteresis and long term stability remain a challenge to PS-BPLC devices and systems. In principle, a higher monomer concentration helps to produce sturdier polymer network which is favorable to suppress hysteresis, but the tradeoff is increased driving voltage [105]. Nevertheless, there has been limited investigation on the impact of UV curing condition upon polymer morphology and hysteresis [106-109]. Besides, most BPLC precursors contain some photo-initiator in order to facilitate the polymerization process [38,43]. However, photo-initiator introduces additional ions to the BPLC system, which not only lowers the resistivity but also causes image sticking.

Therefore, it is important to investigate whether we can produce a high performance PS-BPLC without photoinitiator.

In this chapter, we investigate the UV curing process of BPLC by varying the exposure wavelength and dosage with and without a photo-initiator. We find that through optimizing the

UV exposure conditions, it is possible to fabricate a hysteresis-free, high-stability PS-BPLC device without using photoinitiator. There is no doubt that our result would facilitate the commercialization of BPLC device in the display industry.

68 4.1 PS-BPLC Cured at λ=385 nm

The recipe of the BPLC precursor used in our experiment is listed as follows: LC host

84.10 wt% HTG-135200-100 (HCCH, China), chiral dopant 4.00% R5011 (Sigma Aldrich), and two monomers: 7.11% RM257 (HCCH, China) and 4.79% C12A (Sigma Aldrich). In this study, we did not use any photoinitiator; we will show some results with photoinitiator in Section 4.3.

The precursor was filled into an IPS (in-plane-switching) cell, which has no surface rubbing treatment. In an IPS cell, the top substrate has no electrode, but bottom substrate has interdigitated stripe ITO (indium tin oxide) electrodes: the electrode width is 8μm and electrode gap is 12μm. The cell gap was controlled at 7.5μm by spacer balls.

In experiments, we prepared two identical BPLC cells but exposed UV light from different sides, as Figure 4-1 shows. Sample 1 was exposed from the top substrate side (no electrode), whereas sample 2 was from the bottom side (with IPS electrodes). Two UV wavelengths were used in our studies: λ=365nm (Mercury lamp) and λ=385nm (UV LED). For convenience, let us abbreviate the shorter UV wavelength (λ=365nm) as SUV and longer UV

(λ=385nm) as LUV. The results shown in Figure 4-1 were conducted at LUV with intensity

~8mW/cm2 and exposure time ~40min.

69

Figure 4-1. (Color online) Two UV exposure directions: sample 1 is exposed from top substrate side (no electrode) and sample 2 is from bottom side (IPS electrode). LUV λ=385nm

Next, we measured the voltage-dependent transmittance (VT) curves of both samples.

The IPS cell was sandwiched between two crossed polarizers, and driven with 1 kHz AC voltage signal. A He-Ne laser (λ=633nm) was used as probing beam. Figure 4-2 shows the measured results.

From Figure 4-2, the peak voltages of sample 1 and sample 2 are almost the same

(109.4Vrms vs. 110Vrms). By fitting these VT curves using extended Kerr model, we obtained the

Kerr constant K~2.9nm/V2 for both Samples 1 and 2 [36]. However, Sample 1 has a smaller hysteresis than Sample 2 (0.3% vs. 1.8%), even though both samples are fully cured. This difference, although not significant, could originate from the non-uniform polymer network of sample 2. In sample 2, the ITO electrode blocks a portion of the incident UV light before it reaches the BPLC layer. Therefore, the polymer network in sample 1 will be more uniform and more rigid than in sample 2, which results in a smaller hysteresis for sample 1.

70

Figure 4-2. (Color online) Measured VT curves of samples 1~4: Blue dashed lines are for sample 1, red dotted lines are for sample 2, black solid lines are for sample 3, and magenta dash-and-dot lines are for sample 4.

We also prepared some similar samples but cured at a weaker UV dosage, such as

4mW/cm2 for 40 minutes and 8mW/cm2 for 10 minutes. Our tested results indicate that these samples were not fully cured. Thus for the LUV light source, we need a higher dosage

(8mW/cm2 for 40min) to fully cure the IPS BPLC samples.

4.2 PS-BPLC Cured by λ=365 nm

It has been reported before that the wavelength of a UV light source affects the electro- optical property of a PS-BPLC [106]. So in this Section, we compare the curing effect with SUV

71 (λ=365nm) from a mercury lamp. Again, we prepared two identical IPS BPLC cells: samples 3 and 4. The UV dosage was controlled at 8mW/cm2 for 10min. Sample 3 was exposed from top side, whereas sample 4 from bottom side; similar to the curing directions of sample 1 and sample

2, respectively. The measured VT curves of samples 3 and 4 are also included in Figure 4-2.

From Figure 4-2, we notice that different UV exposure directions result in remarkably different V-T characteristics of samples 3 and 4, although both samples are fully cured. Sample 3 has a much lower peak voltage (88.8Vrms) than sample 4 (118Vrms). The fitted Kerr constant of sample 3 and sample 4 is K=4.46nm/V2 and 2.52nm/V2, respectively. Besides, sample 3 shows a much larger hysteresis than sample 4 (2.9% vs. 0.8%).

Please note that samples 3 and 4 were fully cured under 8mW/cm2 for 10 minutes. This dosage is only 25% of that we employed to cure samples 1 and 2 using λ=385nm. Such a dramatically different behavior is believed to originate from the UV penetration depth inside the

BPLC layer.

The absorption of a liquid crystal composition is determined by the conjugation length of each individual component [110]. We measured the transmission spectrum of our BPLC precursor. In experiment, we made a LC cell (cell gap=22μm) using two UV transparent BaF2 substrates. No ITO or alignment layer was used. We first heated the sample to an isotropic state and measured its transmission spectrum. Next, we cooled the sample to BP-I state and measured its transmission spectrum again. The measured results are normalized to the transmission spectrum of an empty BaF2 cell, as shown in Figure 4-3.

72

Figure 4-3. (Color online) Measured transmission spectrum of BPLC precursor in an isotropic state (blue solid line) and BP-I state (red dashed lines)

In Figure 4-3, the blue line represents the measured data in an isotropic state, where the scattering loss is negligible so that the optical loss is mainly due to the absorption of BPLC precursor. In the visible spectral region, the BPLC sample is highly transparent. As the wavelength decreases to below ~370nm, the transmittance declines (i.e., absorption increases) sharply. The red dashed lines in Figure 4-3 represent the measured data in a BP-I state. Clearly, the Bragg reflection of the BPLC sample appears at λB~460nm (which depends on the chiral concentration) and some light scattering loss is observed. From the blue curve, we find that the absorption coefficient at λ=365nm is ~4X higher than that at λ=385nm. These data explain well why we need 4X higher dosage at λ=385nm in order to fully cure the sample than that cured at

λ=365nm. It is also noticed from Figure 4-3 that the absorption of the precursor in the visible

73 region is smaller than that at 385nm. This is also consistent to the experimental results reported in [108], in which the PS-BPLC cured with a visible light exhibits a more symmetric current- voltage (I-V) curve than that cured with a UV light.

As Figure 4-3 shows, the employed BPLC precursor has much stronger absorption at

λ=365nm than that at 385nm. As a result, the SUV light intensity decays exponentially as it traverses the BPLC cell and the polymer network distribution inside the cell will not be uniform

[111]. Upon UV exposure, more monomers will be attracted to the entrance side and these monomers will crosslink more quickly. Thus, the polymer network concentration will be denser near the top substrate, as Figure 4-4 illustrates. On the other hand, LUV can penetrate the BPLC cell more deeply because of its weaker absorption. As a result, the polymer network is more uniform.

Figure 4-4. (Color online) Polymerization of BPLC cells under different UV exposure conditions

As shown in Figure 4-4, sample 3 was cured with its top substrate (no electrode) facing

SUV. Thus, a denser and more rigid polymer network is formed near the top substrate, whereas

74 the polymer network near the IPS electrodes is looser. On the contrary, sample 4 has denser polymer network near the IPS electrode, as Figure 4-4 depicts. Moreover, the electric field in an

IPS cell is mainly concentrated near the electrodes and its penetration depth depends on the electrode dimension [46]. Based on these mechanisms, sample 3 should have a lower operation voltage but larger hysteresis than sample 4 [107]. Indeed, these phenomena are observed experimentally. In the case of LUV exposure, less UV energy is absorbed by the monomers, so a higher dosage is required to stabilize the BPLC samples. The merit is that the LUV light intensity is more uniform across the LC cell and so does the polymer network. Therefore, no matter from which direction the cell is cured, the driving voltage remains more or less the same

(samples 1 & 2).

4.3 Impact of Photoinitiator

A photoinitiator is a chemical compound that decomposes into free radicals when exposed to UV light. These free radicals play an important role to promote polymerization reactions. Therefore, it helps lower the UV exposure dosage. But on the downside, it introduces extra ions to the LC system, which could cause unfavorable image sticking problem. Therefore, the amount of photoinitiator should be controlled carefully.

In the abovementioned BPLC samples 1-4, we did not use any photoinitiator. In this

Section, we report the impact of photoinitiator in the polymerization process of BPLC. In our experiment, we mixed a small amount of photoinitiator Darocur 1173 (Ciba) to our BPLC precursor mentioned in Section 4.1 and then exposed different UV lights to cure the precursor.

We prepared two samples: BPLC-1 containing 99.8wt% precursor and 0.2wt% photoinitiator,

75 and BPLC-2 containing 97.9wt% precursor and 2.1wt% photoinitiator. Samples 5 and 6 are the

IPS cells filled with BPLC-1 and BPLC-2, respectively, and cured under the same conditions:

λ=365nm, 8mW/cm2 intensity, top-side exposure for 10min.

Figure 4-5 depicts the measured VT curves of sample 3 (no photoinitiator), sample 5

(0.2%) and sample 6 (2.1%) under same UV exposure conditions. As the photoinitiator concentration increases, the hysteresis varies from 2.9% (sample 1) to 1.4% (sample 5) and 1.7%

(sample 6), but the peak voltage increases from 90Vrms to ~95Vrms. These results can be explained as follows. The photoinitiator promotes the polymerization process so that even a low intensity UV near the bottom of the cell can still crosslink the monomers. Thus for the samples exposed from top substrate, photoinitiator helps increase the polymer concentration near the bottom electrodes, which in turn leads to a higher driving voltage but a lower hysteresis.

76

Figure 4-5. (Color online) Measured VT curves of three PS-BPLC samples: (a) black solid line is for sample 3 (no photoinitiator); (b) blue dashed lines are for sample 5 (0.2% photoinitiator); (c) red dotted lines are for sample 6 (2.1% photoinitiator). UV lamp: λ=365nm, top exposure. Probing wavelength λ=633nm

We also cured an IPS cell filled with BPLC-1 (0.2% photoinitiator) under LUV (385nm).

The dosage was 8mW/cm2 x 10min. This dosage is only ¼ of the previous dosage we used to cure samples 1 and 2. With the help of photoinitiator, such a low dosage can still stabilize the

BPLC. The on-state voltage of this sample is similar to that of samples 1 and 2. This result confirms that a small amount of photoinitiator indeed speeds up the curing process of BPLC precursor at λ=385nm.

77 In summary, we find that LUV not only improves polymer network uniformity but also suppresses hysteresis. But to fully cure the BPLC precursor, a higher exposure dosage is needed.

Photoinitiator improves the curing speed for LUV and polymer network uniformity for SUV. But for LUV, the improvement of curing speed may not be worthwhile because photoinitiator introduces extra ions to the LC system, which lowers the resistivity and causes image sticking.

Therefore, from device performance viewpoint following curing conditions are preferred: 1) to employ a LUV, 2) to expose light from top substrate (no electrode), and 3) to avoid using photoinitiator while tolerating the increased exposure dosage.

4.4 Hysteresis-free PS-BPLC

Based on the approaches discussed above, we prepared another BPLC sample with optimal curing conditions. We filled the precursor without photoinitiator into an IPS cell and cured by a 385nm UV light from top side. The UV intensity was 16mW/cm2 and exposure time was 40min. An aging test was taken to evaluate the long term stability of this sample. We applied the peak driving voltage (~115V) upon the same sample for 10 minutes, 20 minutes, 30 minutes, 24 hours and 120 hours, respectively, and then measured its VT curve after each test.

The measured results are plotted in Figure 4-6: the peak voltage of this sample has a small variation of 2.5%. The hysteresis also varies slightly, but is within 1%. We also notice that the residual transmittance after aging test remains at 0.05%, corresponding to CR~1200:1. These results prove that this PS-BPLC sample has quite good stability. Although the driving voltage of this sample is still high compared to other reported BPLC materials, our main objective here is to

78 demonstrate a polymerization process which can lead to hysteresis-free device with good long term stability, which is critical for the commercialization of BPLC in display industry.

Figure 4-6. (Color online) Aging test of a PS-BPLC sample cured under optimized curing conditions. Blue open squares show the hysteresis variation during the test. Red open circles represent the peak voltage variation of the sample during the test.

79 CHAPTER 5 SUMMARY

In this dissertation, we investigate the possibility of using polymer-stabilized blue phase liquid crystal in the design of autostereoscopic 2D/3D switchable display systems.

In Chapter II, we analyze the optical rotatory power of blue phase liquid crystal material, proving that this property is the primary cause for lower contrast ratio of current display systems based on PS-BPLC material. And then we propose several different methods to compensate the optical rotatory power of BPLC and to improve the contrast ratio of BPLC display. The first design (rotating analyzer) is simple and cheap, but the improvement of contrast is also limited.

The last proposal based on dispersive +A film is complicated and expensive, but it is capable of reaching the theoretic limit of contrast ratio determined by the polarizer parameters.

Then in Chapter III, several different photonic devices based on PS-BPLC are proposed for 2D/3D switchable display systems, including two types of adaptive focal length PS-BPLC lenticular microlens array and a high transmittance PS-BPLC slit array. The first lens design is based on two cascaded PS-BPLC layers so as to reduce driving voltage. The fabrication process of such lens, however, is quite challenging. The second lens design simplifies the fabrication process by using resistive film as the top electrode, controlling electric field distribution inside the liquid crystal cell. A 2D/3D display system based on these two kinds of BPLC lens arrays has high optical efficiency and high switching speed between 2D and 3D modes. But due to the high driving voltage of a BPLC lens array, the power consumption of the display system may be quite significant. Therefore, another design based on switchable BPLC slit array is also discussed. Due to the fast response time of PS-BPLC devices, the 3D display system achieves higher dynamic

80 resolution than conventional ones based on fixed parallax barrier. A high transmittance BPLC slit design is also simulated to ensure high over-all optical efficiency of the 3D display system.

Chapter IV addresses another important problem of BPLC devices and systems: the stability issue. It is proven by experiment that UV exposure condition affects the hysteresis of

PS-BPLC devices. Optimized UV exposure condition, including long wavelength UV light source, uniform UV intensity, sufficient exposure dosage and no photo-initiator, results in a hysteresis-free PS-BPLC device with very good long-term stability, which is quite critical for the commercialization of PS-BPLC devices and systems.

Based on these discussions, we prove that the autostereoscopic 2D/3D switchable display system design based on blue phase liquid crystal has many attractive features and it is a competitive candidate for the next generation display systems.

81 APPENDIX: LIST OF PUBLICATIONS

82 Journal Publications:

1. Y.F. Lan, Y. Liu, P.J. Huang, D. Xu, C. Y. Tsai, C. H. Lin, and S.-T. Wu, “Non-ideal optical isotropy of blue phase liquid crystal,” Appl. Phys. Lett. (Accepted, 2014).

2. S. Xu, Y. Li, Y. Liu, J. Sun, H. Ren, and S. T. Wu, “Fast-response liquid crystal microlens,” Micromachines 5, 300-324 (2014).

3. Y. Liu, S. Xu, D. Xu, J. Yan, Y. Gao, and S.-T. Wu, “A hysteresis-free polymer- stabilized blue phase liquid crystal,” Liq. Cryst. (Accepted, May. 2014).

4. Y. Liu, Y. F. Lan, Q. Hong, and S.-T. Wu, “Compensation film designs for high contrast wide-view blue phase liquid crystal displays,” J. Display Technol. 10, 3-6 (Jan., 2014).

5. H. Ren, S. Xu, Y. Liu, and S.-T. Wu, “Optically anisotropic microlens array film directly formed on a single substrate,” Opt. Express 21, 29304-29312 (Dec., 2013).

6. H. Ren, S. Xu, Y. Liu, and S.-T. Wu, “A plano-convex/biconvex microlens array based on self-assembled photocurable polymer droplets,” J. Mater. Chem. C, 1, 7453-7458 (Oct., 2013).

7. D. Xu, Y. Chen, Y. Liu, and S.-T. Wu, “Refraction effect in an in-plane-switching blue phase liquid crystal cell,” Opt. Express 21, 24721-24735 (Oct., 2013).

8. Y. Liu, Y. Li, and S.-T. Wu, “Polarization independent adaptive lens with two different blue phase liquid crystal layers,” Appl. Opt. 52, 3216-3220 (May., 2013).

9. H. Ren, S. Xu, Y. Liu, and S.-T. Wu, “Switchable focus using a polymeric lenticular microlens array and a polarization rotator,” Opt. Express 21, 7916-7925 (Apr., 2013).

10. Y. Liu, Yi-Fen Lan, H. Zhang, R. Zhu, D. Xu, C.Y. Tsai, J.K. Lu, N. Sugiura, Y.C. Lin and S.-T. Wu, “Optical rotatory power of polymer-stabilized blue phase liquid crystals,” Appl. Phys. Lett. 102, 131102 (Apr., 2013).

11. Y. Li, Y. Chen, J. Yan, Y. Liu, J. P. Cui, Q. H. Wang, and S.-T. Wu, “Polymer- stabilized blue phase liquid crystal with a negative Kerr constant,” Opt. Mater. Express 2, 1135-1140 (Aug., 2012).

12. H. Ren, S. Xu, Y. Liu, and S.-T. Wu, “Liquid-based infrared optical switch,” Appl. Phys. Lett. 101, 041104 (Jul., 2012).

13. S. Xu, H. Ren, Y. Liu, and S.-T. Wu, “Color displays based on voltage-stretchable liquid crystal droplet,” J. Display Technol. 8, 336-340 (Jun., 2012).

14. Y. Li, Y. Liu, Q. Li, and S.-T. Wu, “Polarization independent blue-phase cylindrical lens with a resistive film,” Appl. Opt. 51, 2568-2572 (May., 2012).

83 15. M. Xu, H. Ren, C. Nah, S. H. Lee, and Y. Liu, “Liquid crystal micro-lenticular array assembled by a fringing field,” J. Appl. Phys. 111, 063104 (Mar., 2012).

16. Y. Liu, H. Ren, S. Xu, Y. Chen, L. Rao, T. Ishinabe, and S.-T. Wu, “Adaptive focus integral image system design based on Fast-response liquid crystal microlens,” J. of Disp. Tech. 7, 674-678 (Dec. 2011).

17. H. C. Cheng, S. Xu, Y. Liu, S. Levi, and S.-T. Wu, “Adaptive mechanical-wetting lens actuated by ferrofluids,” Opt. Commuin. 284, 2118-2121 (Apr., 2011).

18. H. Ren, S. Xu, Y. Liu, and S.-T. Wu, “Electro-optical properties of dielectric liquid microlens,” Opt. Commuin. 284, 2122-2125 (Apr., 2011).

19. S. Xu, H. Ren, Y. Liu, and S.-T. Wu, “Dielectric liquid microlens with switchable negative and positive optical power,” IEEE J.MEMS 20, 297-301 (Feb., 2011).

20. S. Xu, Y. Liu, H. Ren and S.-T. Wu, “A novel adaptive mechanical-wetting lens for visible and near infrared imaging”, Opt. Express 18, 12430-12435 (Jun., 2010).

Conference Proceedings:

1. Y. Liu, S. Xu, D. Xu, J. Yan, and S.-T. Wu, “A hysteresis-free polymer-stabilized blue phase liquid crystal,” SID Symposium Digest, 45, (Accepted) (San Diego, CA 2014).

2. D. Xu, Y. Chen, Y. Liu, S.-T. Wu “Low-Voltage and High-Transmittance Blue Phase Liquid Crystal Displays,” SID Symposium Digest, 45, (Accepted) (San Diego, CA 2014).

3. Y. Liu, H. Ren, S. Xu, Y. Li, S.-T. Wu, “Fast-response liquid crystal lens for 3D displays,” SPIE Photonics West, 9005, 900503 (San Francisco, CA 2014).

4. Y. Liu, H. Zhang, R. Zhu, D. Xu, and S.-T. Wu, Y.-F. Lan, C.-Y. Tsai, J.-K. Lu, N. Sugiura, and Y.-C. Lin, “Enhancing the Contrast Ratio of Blue Phase LCDs,” SID Symposium Digest, 44, 188-191 (Vancouver, Canada 2013).

5. Y. Li, Y. Liu, and S.-T. Wu, “A Polarization Independent Microlens using Two Blue- Phase Liquid Crystal Layers with Different Kerr Constant,” SID Symposium Digest, 44, 1282-1285 (Vancouver, Canada 2013).

6. Y. Liu, Y. Li, D. Lai, J.-W. Shiu, and S.-T. Wu, “High Transmittance Blue-phase LCD with a Floating Electrode,” SID Symposium Digest, 44, 1279-1281 (Vancouver, Canada 2013).

7. Y. Liu, Y. Li, C.-T. Lee, H.-Y. Lin, Q. Li, and S.-T. Wu, “Polarization-independent and Fast-response Blue Phase Liquid Crystal Lens with a PEDOT: PSS Film,” SID Symposium Digest, 43, 205-207 (Boston, MA 2012).

84 8. S. Xu, H. Ren, Y. Liu, and S.-T. Wu, “Color Displays Based On Voltage-Stretchable Liquid Crystal Droplet,” SID Symposium Digest, 43, 621-624 (Boston, MA 2012).

9. Y. Liu, H. Ren, S. Xu, L. Rao, and S.-T. Wu, “Adaptive Liquid Crystal Lens for Integral Image 3D Display,” SID Symposium Digest, 42, 1596-1598 (Los Angeles, CA 2011).

85 LIST OF REFERENCES

1. S. Liu and H. Hua, "Time-Multiplexed Dual-Focal Plane Head-Mounted Display with a Liquid Lens", Opt. Lett. 34 (11), 1642-1644, (2009).

2. A. K. Srivastava, J. L. de Bougrenet de la Tocnaye and L. Dupont, "Liquid Crystal Active Glasses for 3d Cinema", J. Display Technol. 6 (10), 522-530, (2010).

3. S. Pastoor and M. Wöpking, "3-D Displays: A Review of Current Technologies", Displays 17 (2), 100-110, (1997).

4. G. D. Love, D. M. Hoffman, P. J. W. Hands, J. Gao, A. K. Kirby and M. S. Banks, "High-Speed Switchable Lens Enables the Development of a Volumetric Stereoscopic Display", Opt. Express 17 (18), 15716-15725, (2009).

5. S. C. McQuaide, E. J. Seibel, R. Burstein and T. A. Furness, "50.4: Three-Dimensional System Using a Deformable Membrane Mirror", SID Symposium Digest of Technical Papers 33 (1), 1324-1327, (2002).

6. A. Shiraki, N. Takada, M. Niwa, Y. Ichihashi, T. Shimobaba, N. Masuda and T. Ito, "Simplified Electroholographic Color Reconstruction System Using Graphics Processing Unit and Liquid Crystal Display Projector", Opt. Express 17 (18), 16038-16045, (2009).

7. M. G. H. Hiddink, S. T. de Zwart, O. H. Willemsen and T. Dekker, "20.1: Locally Switchable 3d Displays", SID Symposium Digest of Technical Papers 37 (1), 1142-1145, (2006).

8. S. Sakamoto and Y. Takaki, "Elimination of Flipped Image and Enhancement of Viewing Angle for Lenticular 3d Display", Proc. SPIE 6778, 67780D, (2007).

9. O. H. Willemsen, S. T. de Zwart, M. G. H. Hiddink, D. K. G. de Boer and M. P. C. M. Krijn, "28.1: Invited Paper: Multi-View 3d Displays", SID Symposium Digest of Technical Papers 38 (1), 1154-1157, (2007).

10. D. K. G. de Boer, M. G. H. Hiddink, M. Sluijter, O. H. Willemsen and S. T. de Zwart, "Switchable Lenticular Based 2d/3d Displays", Proc. SPIE 6490, 64900R, (2007).

11. H. Kim, J. Hahn and H.-J. Choi, "Numerical Investigation on the Viewing Angle of a Lenticular Three-Dimensional Display with a Triplet Lens Array", Appl. Opt. 50 (11), 1534-1540, (2011).

12. Q.-H. Wang, X.-F. Li, L. Zhou, A.-H. Wang and D.-H. Li, "Cross-Talk Reduction by Correcting the Subpixel Position in a Multiview Autostereoscopic Three-Dimensional Display Based on a Lenticular Sheet", Appl. Opt. 50 (7), B1-B5, (2011).

86 13. H. J. Lee, H. Nam, J. D. Lee, H. W. Jang, M. S. Song, B. S. Kim, J. S. Gu, C. Y. Park and K. H. Choi, "8.2: A High Resolution Autostereoscopic Display Employing a Time Division Parallax Barrier", SID Symposium Digest of Technical Papers 37 (1), 81-84, (2006).

14. Y. Kim, K. Hong, J. Yeom, J. Hong, J.-H. Jung, Y. W. Lee, J.-H. Park and B. Lee, "A Frontal Projection-Type Three-Dimensional Display", Opt. Express 20 (18), 20130- 20138, (2012).

15. J. Yan, S. T. Kowel, H. J. Cho and C. H. Ahn, "Real-Time Full-Color Three-Dimensional Display with a Micromirror Array", Opt. Lett. 26 (14), 1075-1077, (2001).

16. T. Toda, S. Takahashi and F. Iwata, "Three-dimensional (3D) video system using grating image", Proc. SPIE 2406, 191-198, (1995).

17. J.-H. Park, S. Jung, H. Choi and B. Lee, " with Multiple Image Planes Using a Uniaxial Crystal Plate", Opt. Express 11 (16), 1862-1875, (2003).

18. A. Stern and B. Javidi, "Three-Dimensional Image Sensing, Visualization, and Processing Using Integral Imaging", Proc. IEEE 94 (3), 591-607, (2006).

19. R. Martínez-Cuenca, H. Navarro, G. Saavedra, B. Javidi and M. Martinez-Corral, "Enhanced Viewing-Angle Integral Imaging by Multiple-Axis Telecentric Relay System", Opt. express 15 (24), 16255-16260, (2007).

20. G. Baasantseren, J.-H. Park, K.-C. Kwon and N. Kim, "Viewing Angle Enhanced Integral Imaging Display Using Two Elemental Image Masks", Opt. Express 17 (16), 14405- 14417, (2009).

21. A. Tolosa, R. Martínez-Cuenca, A. Pons, G. Saavedra, M. Martínez-Corral and B. Javidi, "Optical Implementation of Micro-Zoom Arrays for Parallel Focusing in Integral Imaging", J. Opt. Soc. Am. A 27 (3), 495-500, (2010).

22. C. Myungjin, M. Daneshpanah, M. Inkyu and B. Javidi, "Three-Dimensional Optical Sensing and Visualization Using Integral Imaging", Proc. IEEE 99 (4), 556-575, (2011).

23. S.-W. Choi, S.-I. Yamamoto, Y. Haseba, H. Higuchi and H. Kikuchi, "Optically Isotropic-Nanostructured Liquid Crystal Composite with High Kerr Constant", Appl. Phys. Lett. 92 (4), 043119, (2008).

24. H. Kikuchi, M. Yokota, Y. Hisakado, H. Yang and T. Kajiyama, "Polymer-Stabilized Liquid Crystal Blue Phases", Nat. Mater. 1 (1), 64-68, (2002).

25. J. Yan, L. Rao, M. Jiao, Y. Li, H.-C. Cheng and S.-T. Wu, "Polymer-Stabilized Optically Isotropic Liquid Crystals for Next-Generation Display and Photonics Applications", J. Mater. Chem. 21 (22), 7870-7877, (2011).

87 26. H. Lee, H.-J. Park, O.-J. Kwon, S. J. Yun, J. H. Park, S. Hong and S.-T. Shin, "11.1: Invited Paper: The World's First Blue Phase Liquid Crystal Display", SID Symposium Digest of Technical Papers 42 (1), 121-124, (2011).

27. L. Rao, J. Yan, S.-T. Wu, S.-i. Yamamoto and Y. Haseba, "A Large Kerr Constant Polymer-Stabilized Blue Phase Liquid Crystal", Appl. Phys. Lett. 98 (8), 081109, (2011).

28. M. Wittek, N. Tanaka, D. Wilkes, M. Bremer, D. Pauluth, J. Canisius, A. Yeh, R. Yan, K. Skjonnemand and M. Klasen-Memmer, "4.4: New Materials for Polymer-Stabilized Blue Phase", SID Symposium Digest of Technical Papers 43 (1), 25-28, (2012).

29. K.-M. Chen, S. Gauza, H. Xianyu and S.-T. Wu, "Hysteresis Effects in Blue-Phase Liquid Crystals", J. Display Technol. 6 (8), 318-322, (2010).

30. L. Rao, J. Yan, S.-T. Wu, Y.-C. Lai, Y. H. Chiu, H.-Y. Chen, C.-C. Liang, C.-M. Wu, P.- J. Hsieh, S.-H. Liu and K.-L. Cheng, "Critical Field for a Hysteresis-Free BPLC Device", J. Display Technol. 7 (12), 627-629, (2011).

31. D. Xu, Y. Chen, Y. Liu and S.-T. Wu, "Refraction Effect in an in-Plane-Switching Blue Phase Liquid Crystal Cell", Opt. Express 21 (21), 24721-24735, (2013).

32. J. Yan, S.-T. Wu, K.-L. Cheng and J.-W. Shiu, "A Full-Color Reflective Display Using Polymer-Stabilized Blue Phase Liquid Crystal", Appl. Phys. Lett. 102 (8), 081102, (2013).

33. Y. Chen and S.-T. Wu, "Electric Field-Induced Monodomain Blue Phase Liquid Crystals", Appl. Phys. Lett. 102 (17), 171110, (2013).

34. J. Yan, Y. Li and S.-T. Wu, "High-Efficiency and Fast-Response Tunable Phase Grating Using a Blue Phase Liquid Crystal", Opt. Lett. 36 (8), 1404-1406, (2011).

35. Z. Ge, S. Gauza, M. Jiao, H. Xianyu and S.-T. Wu, "Electro-Optics of Polymer-Stabilized Blue Phase Liquid Crystal Displays", Appl. Phys. Lett. 94 (10), 101104, (2009).

36. J. Yan, H.-C. Cheng, S. Gauza, Y. Li, M. Jiao, L. Rao and S.-T. Wu, "Extended Kerr Effect of Polymer-Stabilized Blue-Phase Liquid Crystals", Appl. Phys. Lett. 96 (7), 071105, (2010).

37. K.-M. Chen, S. Gauza, H. Xianyu and S.-T. Wu, "Submillisecond Gray-Level Response Time of a Polymer-Stabilized Blue-Phase Liquid Crystal", J. Display Technol. 6 (2), 49- 51, (2010).

38. Y. Chen, J. Yan, J. Sun, S.-T. Wu, X. Liang, S.-H. Liu, P.-J. Hsieh, K.-L. Cheng and J.- W. Shiu, "A Microsecond-Response Polymer-Stabilized Blue Phase Liquid Crystal", Appl. Phys. Lett. 99 (20), 201105, (2011).

88 39. H.-C. Cheng, J. Yan, T. Ishinabe and S.-T. Wu, "Vertical Field Switching for Blue-Phase Liquid Crystal Devices", Appl. Phys. Lett. 98 (26), 261102, (2011).

40. H.-C. Cheng, J. Yan, T. Ishinabe, C.-H. Lin, K.-H. Liu and S.-T. Wu, "Wide-View Vertical Field Switching Blue-Phase LCD", J. Display Technol. 8 (11), 627-633, (2012).

41. H.-C. Cheng, J. Yan, T. Ishinabe, N. Sugiura, C.-Y. Liu, T.-H. Huang, C.-Y. Tsai, C.-H. Lin and S.-T. Wu, "Blue-Phase Liquid Crystal Displays with Vertical Field Switching", J. Display Technol. 8 (2), 98-103, (2012).

42. L. Rao, H.-C. Cheng and S.-T. Wu, "Low Voltage Blue-Phase LCDs with Double- Penetrating Fringe Fields", J. Display Technol. 6 (8), 287-289, (2010).

43. Y. Chen, D. Xu, S.-T. Wu, S.-I. Yamamoto and Y. Haseba, "A Low Voltage and Submillisecond-Response Polymer-Stabilized Blue Phase Liquid Crystal", Appl. Phys. Lett. 102 (14), 141116, (2013).

44. Y. Haseba, H. Kikuchi, T. Nagamura and T. Kajiyama, "Large Electro-Optic Kerr Effect in Nanostructured Chiral Liquid-Crystal Composites over a Wide Temperature Range", Adv. Mater. 17 (19), 2311-2315, (2005).

45. Y. Hisakado, H. Kikuchi, T. Nagamura and T. Kajiyama, "Large Electro-Optic Kerr Effect in Polymer-Stabilized Liquid-Crystalline Blue Phases", Adv. Mater. 17 (1), 96-98, (2005).

46. L. Rao, Z. Ge, S.-T. Wu and S. H. Lee, "Low Voltage Blue-Phase Liquid Crystal Displays", Appl. Phys. Lett. 95 (23), 231101, (2009).

47. M. Jiao, Y. Li and S.-T. Wu, "Low Voltage and High Transmittance Blue-Phase Liquid Crystal Displays with Corrugated Electrodes", Appl. Phys. Lett. 96 (1), 011102, (2010).

48. Y. Li and S.-T. Wu, "Transmissive and Transflective Blue-Phase LCDs with Enhanced Protrusion Electrodes", J. Display Technol. 7 (7), 359-361, (2011).

49. M. Kim, M. S. Kim, B. G. Kang, M.-K. Kim, S. Yoon, S. H. Lee, Z. Ge, L. Rao, S. Gauza and S.-T. Wu, "Wall-Shaped Electrodes for Reducing the Operation Voltage of Polymer-Stabilized Blue Phase Liquid Crystal Displays", J. Phys. D: Appl. Phys. 42 (23), 235502, (2009).

50. S. Yoon, M. Kim, M. S. Kim, B. G. Kang, M.-K. Kim, A. K. Srivastava, S. H. Lee, Z. Ge, L. Rao, S. Gauza and S.-T. Wu, "Optimisation of Electrode Structure to Improve the Electro-Optic Characteristics of Liquid Crystal Display Based on the Kerr Effect", Liq. Cryst. 37 (2), 201-208, (2010).

51. S. Sato and T. Nose, "Liquid Crystal Microlens and Improvement of the Properties", Proc. SPIE 3800, 72-86, (1999).

89 52. H. Ren, Y.-H. Fan, S. Gauza and S.-T. Wu, "Tunable-Focus Flat Liquid Crystal Spherical Lens", Appl. Phys. Lett. 84 (23), 4789-4791, (2004).

53. H. Ren, Y.-H. Fan, S. Gauza and S.-T. Wu, "Tunable-Focus Cylindrical Liquid Crystal Lens", Japn. J. Appl. Phys. 43 (2), 652-653, (2004).

54. Y.-H. Fan, H. Ren, X. Liang, H. Wang and S.-T. Wu, "Liquid Crystal Microlens Arrays with Switchable Positive and Negative Focal Lengths", J. Display Technol. 1 (1), 151- 156, (2005).

55. M. Ye, B. Wang and S. Sato, "Development of High-Quality Liquid Crystal Lens", Proc. SPIE 6487, 64870N, (2007).

56. H. T. Dai, Y. J. Liu, X. W. Sun and D. Luo, "A Negative-Positive Tunable Liquid-Crystal Microlens Array by Printing", Opt. Express 17 (6), 4317-4323, (2009).

57. M. Honma, T. Nose, S. Yanase, R. Yamaguchi and S. Sato, "Liquid-Crystal Variable- Focus Lenses with a Spatially-distributed Tilt Angles", Opt. Express 17 (13), 10998- 11006, (2009).

58. Y.-Y. Kao, P. C. P. Chao and C.-W. Hsueh, "A New Low-Voltage-Driven Grin Liquid Crystal Lens with Multiple Ring Electrodes in Unequal Widths", Opt. Express 18 (18), 18506-18518, (2010).

59. K. Asatryan, V. Presnyakov, A. Tork, A. Zohrabyan, A. Bagramyan and T. Galstian, "Optical Lens with Electrically Variable Focus Using an Optically Hidden Dielectric Structure", Opt. Express 18 (13), 13981-13992, (2010).

60. A. Y. G. Fuh, S.-W. Ko, S.-H. Huang, Y.-Y. Chen and T.-H. Lin, "Polarization- Independent Liquid Crystal Lens Based on Axially Symmetric Photoalignment", Opt. Express 19 (3), 2294-2300, (2011).

61. C.-H. Lin, Y.-Y. Wang and C.-W. Hsieh, "Polarization-Independent and High- Diffraction-Efficiency Fresnel Lenses Based on Blue Phase Liquid Crystals", Opt. Lett. 36 (4), 502-504, (2011).

62. H. Hong, "Analysis of Polarisation Change in an Electric Field-Driven Liquid Crystal Lens of Cylindrical Type Where LC Are Aligned Twisted", Liq. Cryst. 38 (6), 689-696, (2011).

63. Y.-H. Lin, H.-S. Chen, H.-C. Lin, Y.-S. Tsou, H.-K. Hsu and W.-Y. Li, "Polarizer-Free and Fast Response Microlens Arrays Using Polymer-Stabilized Blue Phase Liquid Crystals", Appl. Phys. Lett. 96 (11), 113505, (2010).

64. Y. Li and S.-T. Wu, "Polarization Independent Adaptive Microlens with a Blue-Phase Liquid Crystal", Opt. Express 19 (9), 8045-8050, (2011).

90 65. C.-T. Lee, Y. Li, H.-Y. Lin and S.-T. Wu, "Design of Polarization-Insensitive Multi- Electrode Grin Lens with a Blue-Phase Liquid Crystal", Opt. Express 19 (18), 17402- 17407, (2011).

66. Y. Li, Y. Liu, Q. Li and S.-T. Wu, "Polarization Independent Blue-Phase Liquid Crystal Cylindrical Lens with a Resistive Film", Appl. Opt. 51 (14), 2568-2572, (2012).

67. Y. Hirakata, D. Kubota, A. Yamashita, T. Ishitani, T. Nishi, H. Miyake, H. Miyairi, J. Koyama, S. Yamazaki, T. Cho and M. Sakakura, "A Novel Field-Sequential Blue-Phase- Mode AMLCD", J. Soc. Info. Display 20 (1), 38-46, (2012).

68. S. Yoon, G. H. Yang, P. Nayek, H. Jeong, S. H. Lee, S. H. Hong, H. J. Lee and S. T. Shin, "Study on the Light Leakage Mechanism of a Blue Phase Liquid Crystal Cell with Oblique Interfaces", J. Phys. D: Appl. Phys. 45 (10), 105304, (2012).

69. D. Bensimon, E. Domany and S. Shtrikman, "Optical Activity of Cholesteric Liquid Crystals in the Pretransitional Regime and in the Blue Phase", Phys. Rev. A 28 (1), 427- 433, (1983).

70. D. Bensimon, E. Domany and S. Shtrikman, "Optical Activity of Cholesteric Liquid Crystals in the Pretransitional Regime and in the Blue Phase", Phys. Rev. A 28 (1), 427- 433, (1983).

71. Z. Kutnjak, C. W. Garland, C. G. Schatz, P. J. Collings, C. J. Booth and J. W. Goodby, "Critical Point for the Blue-Phase-Iii–Isotropic Phase Transition in Chiral Liquid Crystals", Phys. Rev. E 53 (5), 4955-4963, (1996).

72. B.-Y. Zhang, F.-B. Meng and Y.-H. Cong, "Optical Characterization of Polymer Liquid Crystal Cell Exhibiting Polymer Blue Phases", Opt. Express 15 (16), 10175-10181, (2007).

73. S. Meiboom, J. P. Sethna, P. W. Anderson and W. F. Brinkman, "Theory of the Blue Phase of Cholesteric Liquid Crystals", Phys. Rev. Lett. 46 (18), 1216-1219, (1981).

74. M. Schadt and W. Helfrich, "Voltage-Dependent Optical Activity of a Twisted Nematic Liquid Crystal", Appl. Phys. Lett. 18 (4), 127-128, (1971).

75. J. Yan and S.-T. Wu, "Polymer-Stabilized Blue Phase Liquid Crystals: A Tutorial [Invited]", Opt. Mater. Express 1 (8), 1527-1535, (2011).

76. J. Yan, Y. Chen, S.-T. Wu and X. Song, "Figure of Merit of Polymer-Stabilized Blue Phase Liquid Crystals", J. Display Technol. 9 (1), 24-29, (2013).

77. S.-T. Wu and C.-S. Wu, "Experimental Confirmation of the Osipov-Terentjev Theory on the Viscosity of Nematic Liquid Crystals", Phys. Rev. A 42 (4), 2219-2227, (1990).

91 78. H. Baessler, T. M. Laronge and M. M. Labes, "Electric Field Effects on the Optical Rotatory Power of a Compensated Cholesteric Liquid Crystal", J. Chem. Phys. 51 (8), 3213-3219, (1969).

79. Y. Liu, Y.-F. Lan, H. Zhang, R. Zhu, D. Xu, C.-Y. Tsai, J.-K. Lu, N. Sugiura, Y.-C. Lin and S.-T. Wu, "Optical Rotatory Power of Polymer-Stabilized Blue Phase Liquid Crystals", Appl. Phys. Lett. 102 (13), 131102, (2013).

80. L. Rao, Z. Ge and S.-T. Wu, "Zigzag Electrodes for Suppressing the Color Shift of Kerr Effect-Based Liquid Crystal Displays", J. Display Technol. 6 (4), 115-120, (2010).

81. Q. Hong, T. X. Wu, R. Lu and S.-T. Wu, "Wide-View Circular Polarizer Consisting of a Linear Polarizer and Two Biaxial Films", Opt. Express 13 (26), 10777-10783, (2005).

82. X. Zhu, Z. Ge and S.-T. Wu, "Analytical Solutions for Uniaxial-Film-Compensated Wide-View Liquid Crystal Displays", J. Display Technol. 2 (1), 2-20, (2006).

83. S. Sato, "Liquid-Crystal Lens-Cells with Variable Focal Length", Japn. J. Appl. Phys. 18 (9), 1679-1684, (1979).

84. C.-W. Chen, M. Cho, Y.-P. Huang and B. Javidi, "Three-Dimensional Imaging with Axially Distributed Sensing Using Electronically Controlled Liquid Crystal Lens", Opt. Lett. 37 (19), 4125-4127, (2012).

85. T. Nose, S. Masuda, S. Sato, J. Li, L.-C. Chien and P. J. Bos, "Effects of Low Polymer Content in a Liquid-Crystal Microlens", Opt. Lett. 22 (6), 351-353, (1997).

86. Y.-P. Huang, C.-W. Chen and Y.-C. Huang, "Superzone Fresnel Liquid Crystal Lens for Temporal Scanning Auto-Stereoscopic Display", J. Display Technol. 8 (11), 650-655, (2012).

87. L. Lu, L. Shi, P. J. Bos, T. van Heugten and D. Duston, "14.51: Late-News Paper: Comparisons between a Liquid Crystal Refractive Lens and a Diffractive Lens for 3d Displays", SID Symposium Digest of Technical Papers 42 (1), 171-174, (2011).

88. Y.-P. Huang, C.-W. Chen and T.-C. Shen, "25.1 Invited Paper: High Resolution Autostereoscopic 3d Display with Scanning Multi-Electrode Driving Liquid Crystal (MED-LC) Lens", SID Symposium Digest of Technical Papers 40 (1), 336-339, (2009).

89. M. Ferstl and A.-M. Frisch, "Static and Dynamic Fresnel Zone Lenses for Optical Interconnections", J. Modern Opt. 43 (7), 1451-1462, (1996).

90. P. F. McManamon, T. A. Dorschner, D. L. Corkum, L. J. Friedman, D. S. Hobbs, M. Holz, S. Liberman, H. Q. Nguyen, D. P. Resler, R. C. Sharp and E. A. Watson, "Optical Phased Array Technology", Proc. IEEE 84 (2), 268-298, (1996).

92 91. N. Fraval and J. L. d. B. de la Tocnaye, "Low Aberrations Symmetrical Adaptive Modal Liquid Crystal Lens with Short Focal Lengths", Appl. Opt. 49 (15), 2778-2783, (2010).

92. A. Naumov, G. Love, M. Y. Loktev and F. Vladimirov, "Control Optimization of Spherical Modal Liquid Crystal Lenses", Opt. Express 4 (9), 344-352, (1999).

93. M. Ye and S. Sato, "Optical Properties of Liquid Crystal Lens of Any Size", Japn. J. Appl. Phys. 41 (5B), L571-L573, (2002).

94. N. A. Riza and M. C. DeJule, "Three-Terminal Adaptive Nematic Liquid-Crystal Lens Device", Opt. Lett. 19 (14), 1013-1015, (1994).

95. S. T. Kowel, D. S. Cleverly and P. G. Kornreich, "Focusing by Electrical Modulation of Refraction in a Liquid Crystal Cell", Appl. Opt. 23 (2), 278-289, (1984).

96. B. Wang, M. Ye and S. Sato, "Liquid Crystal Lens with Focal Length Variable from Negative to Positive Values", IEEE Photonics Technol. Lett. 18 (1), 79-81, (2006).

97. H. Ren, D. W. Fox, B. Wu and S.-T. Wu, "Liquid Crystal Lens with Large Focal Length Tunability and Low Operating Voltage", Opt. Express 15 (18), 11328-11335, (2007).

98. S. Gauza, H. Wang, C.-H. Wen, S.-T. Wu, A. J. Seed and R. Dabrowski, "High Birefringence Isothiocyanato Tolane Liquid Crystals", Japn. J. Appl. Phys. 42 (6A), 3463-3466, (2003).

99. Y.-H. Lin, H. Ren, Y.-H. Wu, Y. Zhao, J. Fang, Z. Ge and S.-T. Wu, "Polarization- Independent Liquid Crystal Phase Modulator Using a Thin Polymer-Separated Double- Layered Structure", Opt. Express 13 (22), 8746-8752, (2005).

100. P. R. Gerber, "Electro-Optical Effects of a Small-Pitch Blue-Phase System", Mol. Cryst. Liq. Cryst. 116 (3-4), 197-206, (1985).

101. M. Jiao, J. Yan and S.-T. Wu, "Dispersion Relation on the Kerr Constant of a Polymer- Stabilized Optically Isotropic Liquid Crystal", Phys. Rev. E 83 (4), 041706, (2011).

102. D.-S. Kim, S. Shestak, K.-H. Cha, S.-M. Park and S.-D. Hwang, "Time-Sequential Autostereoscopic OLED Display with Segmented Scanning Parallax Barrier", Proc. SPIE 7329, 73290U, (2009).

103. Y. Chen, Y. Sun and G. Yang, "Low Voltage and High Transmittance Blue-Phase LCDs with Double-Side In-Plane Switching Electrodes", Liq. Cryst. 38 (5), 555-559, (2011).

104. Y. Liu, Y.-F. Lan, Q. Hong and S.-T. Wu, "Compensation Film Designs for High Contrast Wide-View Blue Phase Liquid Crystal Displays", J. Display Technol. 10 (1), 3- 6, (2014).

93 105. J. Yan and S.-T. Wu, "Effect of Polymer Concentration and Composition on Blue Phase Liquid Crystals", J. Display Technol. 7 (9), 490-493, (2011).

106. T. N. Oo, T. Mizunuma, Y. Nagano, H. Ma, Y. Ogawa, Y. Haseba, H. Higuchi, Y. Okumura and H. Kikuchi, "Effects of Monomer/Liquid Crystal Compositions on Electro- Optical Properties of Polymer-Stabilized Blue Phase Liquid Crystal", Opt. Mater. Express 1 (8), 1502-1510, (2011).

107. C.-Y. Fan, C.-T. Wang, T.-H. Lin, F.-C. Yu, T.-H. Huang, C.-Y. Liu and N. Sugiura, "17.3: Hysteresis and Residual Birefringence Free Polymer-Stabilized Blue Phase Liquid Crystal", SID Symposium Digest of Technical Papers 42 (1), 213-215, (2011).

108. S.-I. Yamamoto, T. Iwata, Y. Haseba, D.-U. Cho, S.-W. Choi, H. Higuchi and H. Kikuchi, "Improvement of Electro-Optical Properties on Polymer-Stabilised Optically Isotropic Liquid Crystals", Liq. Cryst. 39 (4), 487-491, (2012).

109. H.-S. Chen, Y.-H. Lin, C.-H. Wu, M. Chen and H.-K. Hsu, "Hysteresis-Free Polymer- Stabilized Blue Phase Liquid Crystals Using Thermal Recycles", Opt. Mater. Express 2 (8), 1149-1155, (2012).

110. S.-T. Wu, E. Ramos and U. Finkenzeller, "Polarized UV Spectroscopy of Conjugated Liquid Crystals", J. Appl. Phys. 68 (1), 78-85, (1990).

111. H. Ren and S.-T. Wu, "Tunable Electronic Lens Using a Gradient Polymer Network Liquid Crystal", Appl. Phys. Lett. 82 (1), 22-24, (2003).

94