throbber
Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`Liquid crystal alignment by rubbed polymer surfaces: a
`microscopic bond orientation model
`J. Sto¨hr*, M.G. Samant
`IBM Research Division, Almaden Research Center, 650 Harry Road, San Jose, CA 95120 USA
`
`Dedication by J. Sto¨hr — This paper is dedicated to Dick Brundle who for many years was my colleage at the IBM Almaden Research Center.
`Dick was responsible for my hiring by IBM, and over the years we interacted with each other in many roles — as each other’s boss or simply as
`colleagues. One thing never changed, we were friends and running buddies. I sure miss those runs with Dick through the Almaden hills and our
`lively discussions. I am sure that he will be missed as an editor of the Journal of Electron Spectroscopy, too!
`
`Received 24 October 1997; accepted 26 December 1997
`
`Abstract
`
`We discuss the microscopic origin of a previously poorly understood phenomenon, the alignment of a nematic liquid crystal
`(LC), consisting of rod-like molecular units, when placed on a rubbed polymer surface. After giving a brief review of the
`phenomenon and its technological utilization in flat panel displays we discuss the use of surface sensitive, polarization
`dependent near edge X-ray absorption spectroscopy for the study of rubbed polymer surfaces. These measurements are
`shown to provide a microscopic picture for the origin of the alignment process. It is shown that the LC orientation direction
`is set by an asymmetry in the molecular bonds, i.e. of the charge, at the rubbed polymer surface. The experimental results are
`explained by a general theory, based on tensor order parameters, which states that the minimum energy state of the interaction
`between the LC and oriented polymer surface corresponds to maximum directional overlap of the respective anisotropic charge
`distributions. q 1998 Elsevier Science B.V. All rights reserved.
`
`Keywords Liquid crystals; Polymer surfaces; Polyimides; Maximum overlap model; Tensor order parameters
`
`1. Introduction
`
`When a nematic liquid crystal (LC), consisting of
`an assembly of aligned rod-like molecules, is placed
`on a rubbed polymer surface it exhibits both in-plane
`and out-of-plane orientation of the rods. The in-plane
`alignment direction of the rods typically coincides
`with the rubbing direction. The average upward tilt
`angle of the rods from the polymer surface plane,
`which is typically a few degrees, is referred to as
`the pretilt angle. It is an important fact that the LC
`
`* Corresponding author. Tel.: 1 1-408-927-2461; Fax: 1 1-
`408-927-2100; e-mail: stohr@almaden.ibm.com
`
`pretilt is unidirectional. For example, for rubbed poly-
`imide surfaces the rods tilt up from the rubbing direc-
`tion and therefore the tipped-up ends of the rods point
`into, not opposite to, the rubbing direction. The origin
`of the LC alignment mechanism has been debated
`ever since its discovery 90 years ago [1,2] but no
`definitive understanding has emerged. Such an under-
`standing is not only an interesting scientific but an
`important
`technological problem since todays flat
`panel displays are largely based on LCs which modu-
`late light transmission from the back to the front of the
`display through changes
`in LC alignment, as
`discussed in more detail below.
`In general, the LC alignment has to originate from
`
`0368-2048/99/$ - see front matter q 1998 Elsevier Science B.V. All rights reserved.
`PII: S0368-2048(98)00286-2
`
`Page 1 of 19
`
`Tianma Exhibit 1016
`
`

`
`190
`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`symmetry breaking at the surface of the polymer
`substrate. Over the years asymmetries in either the
`macroscopic topographical or microscopic molecular
`structure of the polymer surface have been proposed
`for the preferred rod direction in the LC [3,4]. Inde-
`pendent of any asymmetry of the polymer surface, the
`origin and size of out-of-plane LC tilt angles have
`been explained in terms of the van der Waals interac-
`tion between asymmetric LC molecules and the
`polymer surface, modelled as a semi-infinite dielectric
`medium [5]. Because such a model ignores any aniso-
`tropy of the polymer surface on a molecular level it
`cannot account for the unidirectional LC pretilt direc-
`tion, however. While a variety of methods, in parti-
`cular linear and non-linear optical methods, can be
`used to determine the precise alignment direction of
`the LC molecules, even for monolayer films [6], it is
`more difficult to obtain detailed information regarding
`the molecular structure of
`the polymer surface.
`Conventional linear optics techniques lack surface
`sensitivity and second harmonic generation cannot
`be used because the polymer molecules typically
`have inversion symmetry. Nevertheless, early propo-
`sals of the molecular orientation at
`the polymer
`surface were based on bulk-sensitive optical and
`infra-red measurements carried out on thin films [7–
`12].
`More recently, surface sensitive grazing incidence
`X-ray scattering (GIXS) studies on semicrystalline
`BPDA-PDA polyimide demonstrated the preferential
`near-surface alignment of polyimide chain segments
`along the rubbing direction, linking it to the preferred
`in-plane alignment direction of the LCs [13]. These
`studies are consistent with the conventional view that
`LC alignment on the polymer surface originates from
`a quasi-epitaxial
`interaction. Because of the long
`structural coherence length within the LC, the LC
`alignment has been thought to originate from ordered
`regions at the surface with parallel chain segments [7],
`possibly in the form of microcrystalline nucleation
`sites [14]. In this model the LC rods are envisioned
`to align parallel to the polymer chain segments in the
`crystalline regions. Surface sensitive studies have also
`been carried out using the near edge X-ray absorption
`fine structure (NEXAFS) technique [15–19]. These
`studies also showed the preferential near-surface
`alignment of polyimide chain segments along the
`rubbing direction [17] and the decay of the alignment
`
`from the surface toward the bulk of the film. NEXAFS
`studies also gave clear evidence for a preferential out-
`of-plane tilt of phenyl rings at polyimide surfaces
`[15,16,18,19]. This tilt was linked with the pretilt
`angle of the LC on the surface. Some of the NEXAFS
`studies utilized partially ordered (semi-crystalline)
`polyimides like BPDA-PDA [17] and PMDA-ODA
`[18] and the suggested LC alignment models impli-
`citly assumed the presence of ordered surface regions
`giving rise to epitaxial effects. These studies did not
`address the known fact that LC alignment also occurs
`on surfaces of disordered polymers. In fact, such poly-
`mers are typically used in the manufacturing of flat
`panel displays. Most recently NEXAFS studies on a
`disordered polyimide [19] suggested that LC align-
`ment only requires a statistically significant preferen-
`tial bond orientation at the polymer surface, without
`the necessity of crystalline or quasi-crystalline order.
`A general directional interaction model was proposed
`in which the LC direction is guided by a ‘‘p-like
`interaction’’ between the LC molecules and the aniso-
`tropic polymer surface.
`Here we report surface-sensitive and polarization-
`dependent NEXAFS measurements on a variety of
`polymers. The experimentally observed anisotropy
`of the NEXAFS intensity at the polymer surface is
`analyzed in terms of the preferred orientation of
`phenyl and CyO functional groups at the surface.
`The molecular orientation at the polymer surface is
`quantified by the derivation of orientation factors,
`previously utilized for the description of LCs, and
`the relevant equations are derived. From the measure-
`ments a simple general model for LC alignment
`emerges which is based on the existence of preferred
`bond orientation at
`the rubbed polymer surface,
`without the necessity for crystalline or microcrystal-
`line order. In particular, the observed asymmetric out-
`of-plane bond orientation is argued to be the micro-
`scopic origin of the LC pretilt direction. A model is
`presented that links the asymmetric molecular orien-
`tation at the polyimide surface to the rubbing process.
`Finally, we develop a simple but powerful theory for
`the origin of LC alignment that is based on symmetry.
`It uses tensor order parameters to describe the aniso-
`tropic interaction between the LC and the aligned
`polymer surface and clearly shows that the interaction
`energy between a uniaxial LC and a biaxial polymer
`surface is minimized if the p orbital densities of the
`
`Page 2 of 19
`
`

`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`191
`
`Fig. 1. Schematic of a laptop computer flat panel display, as described in the text. On the computer screen we show the structure of a 5 OCB
`liquid crystal molecule.
`
`two systems have maximum directional overlap. This
`theory supports the empirically derived p interaction
`model [19].
`
`2. The role of liquid crystals in flat panel displays
`
`Today’s laptop computers use flat panel displays
`
`Illustration of the rubbing process. A thin polymer film,
`Fig. 2.
`coated on top of an ITO electrode layer which is deposited on a
`glass plate, moves underneath a rotating rubbing wheel whose drum
`is coated with a velour bristle cloth. Also shown is the typical shape
`of a polymer chain, resembling a ball with a radius of gyration of
`about 10 nm and the monomer structure of BPDA-PDA polyimide,
`studied in this paper.
`
`light weight and compact size
`their
`because of
`[20,21]. It is envisioned that such displays will gradu-
`ally replace conventional cathode ray tubes in many
`applications from desk-top computer to television
`screens. In such a display, schematically shown in
`Fig. 1, the picture on the screen is composed of
`many pixels, approximately 300 £ 300 micrometer
`in size, of different colors and intensities [22]. In
`each pixel the desired color is created by ‘‘mixing’’
`blue, green and red primary colors of different inten-
`sities by means of a patterned color filter array as
`shown in Fig. 1. The intensity of each color is adjusted
`by using liquid crystals to change the light intensity
`transmitted from the back to the font of the display.
`The LC is composed of rod like molecules which
`prefer to align themselves so that the long directions
`of the rods are parallel. The structure of a typical LC
`molecule is shown on the computer screen in Fig. 1.
`The LC is filled into the gap, a few microns wide,
`between two polyimide films coated onto indium-
`tin-oxide (ITO) electrodes which in turn are deposited
`onto two glass-plate cross polarizers. In order for the
`display to work the LC molecules have to be anchored
`down nearly parallel to the surfaces of the polyimide
`films but on opposite sides point into the perpendi-
`cular directions of the two crossed polarizers. They
`thus form a twisted helix from one side to the other.
`When the light from a light source in the back of the
`display crosses the first polarizer it is polarized along
`the long axis of the LC molecules anchored to it. As
`
`Page 3 of 19
`
`

`
`192
`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`Fig. 3. Illustration of the liquid crystal pretilt angle 1. For a rubbed polyimide film the pretilt always points into the rubbing direction, indicated
`by a double-stem arrow. In a display, the liquid crystal is filled into the gap between two opposing polyimide coated glass plates which are
`rubbed in orthogonal directions, as shown. The rod-like liquid crystal molecules are anchored down with their long axis parallel to the rubbing
`direction on both ends and form a twisted helix across the gap. Because of the well defined LC pretilt direction at the anchoring points only a
`counterclockwise 908 rotation of the helix is possible when viewed from above, as shown on the right. This avoids the formation of reverse tilt
`domains as discussed in the text.
`
`the light progresses through the LC the helical LC
`structure changes the polarization of the light from
`linear to elliptical so that part of the light is trans-
`mitted by the second, perpendicular, polarizer. Since
`the light transmission depends on the orientation of
`the LC rods it can be changed by rotation of the rods.
`This is accomplished by application of a small
`voltage, pixel by pixel, by means of microscopic
`ITO electrodes independently driven by a transistor
`array. As the voltage is increased the LC long axis
`becomes increasingly parallel
`to the electric field
`direction, which is parallel
`to the light direction.
`The light polarization becomes less affected by the
`LC and the light transmission is reduced because of
`the crossed polarizers. Thus the orientational changes
`in LC alignment are the heart of the LC display
`providing its gray scale or color contrast.
`One of the most important yet scientifically least
`understood steps in the making of a flat panel display
`is the directional anchoring of the LC molecules to the
`polymer films. The current method simply consists of
`unidirectional rubbing of a polyimide film, which is
`about 100 nm thick, with a velour cloth. In practice
`this is done with a rubbing machine, as shown in Fig.
`2, where the polymer coated glass plate moves under-
`neath a rotating drum whose surface consists of velour
`bristles. The rubbing process is specified by the rota-
`tion speed of the drum, the speed of the plate, and the
`‘‘pile impression’’, characterizing the offset
`in
`distance between the drum surface and the polyimide
`surface. Polyimide is chosen because its high glass
`transition temperature assures that the rubbed surface
`
`remains stable even at elevated temperatures. After
`the rubbing process the LC molecules align with
`their long axis parallel to the rubbing direction and
`point up from the surface by a small pretilt angle. The
`size of the pretilt angle depends on the exact monomer
`structure of the polyimide, i.e. it varies with changes
`in main as well as side chain structure, and it depends
`on the structure of the LC molecules, as well.
`The pretilt angle is of particular technological
`importance and, in practice, has to exceed about 38
`for the proper functioning of the display. The reason is
`illustrated in Fig. 3. If the pretilt angle is zero, there
`exists an ambiguity in the twisting of the LC helix
`between the two anchored, orthogonal, ends. Both
`908 clockwise and anticlockwise rotation is possible,
`leading to the formation of clockwise and anticlock-
`wise LC domains, referred to as reverse tilt domains
`[21]. At the boundary between two such domains the
`LC orientation is ill defined leading to artifacts in the
`image on the computer screen, e.g. shadows. In the
`presence of a clearly defined pretilt angle and pretilt
`direction, i.e. into the rubbing direction, only one 908
`twist is possible and therefore the LC will align as a
`single domain. Fig. 3 clearly indicates that a well-
`defined pretilt direction is of fundamental importance.
`For example, if a pretilt angle existed with equal prob-
`ability parallel and antiparallel to the rubbing direc-
`tion, the rotation sense of the helix would remain
`undefined and the reversed domain problem could
`not be eliminated. It is apparent that a detailed under-
`standing of the origin of LC alignment is of great
`technological
`importance. Such an understanding
`
`Page 4 of 19
`
`

`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`193
`
`experimental geometries shown in Figs. 5(a) and
`(b). Here we have chosen our sample coordinate
`system with the x axis along the rubbing direction
`and the z axis along the surface normal. The sample
`could be rotated about a vertical axis resulting in a
`change of the X-ray incidence angle from the surface
`u, and independently about the surface normal z,
`changing the azimuthal angle F of the incident X-
`rays, defined in Fig. 5(c). In the ‘‘parallel’’ geometry
`the major component of the electric field vector, ~E1,
`which lies in the horizontal plane at a 908 angle from
`the incident X-ray wave vector, was oriented in the
`(2x; z; 1x) plane of the sample coordinate system at a
`polar angle u from the sample normal z. For conve-
`nience we define u to be positive for F 08 ~E1 in (1x,
`z) quadrant) and negative for F 1808 ~E1 in (2x, z)
`quadrant). In the ‘‘perpendicular’’ geometry ~E1 was
`oriented in the (2y, z, 1 y) plane at an angle u from
`the sample normal z. Here we define u to be positive
`for F 908 ~E1 in (1y, z) quadrant and negative for
`F 2708 ~E1 in (2y, z) quadrant). X-ray absorption
`was recorded by means of surface sensitive electron
`yield detection [24]. We used a cylindrical mirror
`analyzer (CMA) to monitor the KLL Auger electron
`yield (AEY) which probes only the first nanometer
`from the free surface [17,25]. Simultaneously, we
`measured the sample current with a picoammeter.
`The so obtained total electron yield (TEY) spectra
`probe about 10 nm below the surface [17]. The spectra
`were divided by the total electron yield signal from a
`highly transmissive ( , 80%) gold grid, again
`measured with a picoammeter. A pre-edge back-
`ground was then subtracted from the normalized
`spectra and the edge jump far above the K-edge
`(340–380 eV) was arbitrarily scaled to unity. This
`procedure produces spectra in which all resonance
`intensities correspond to the same number of C or O
`atoms in the sample, as discussed elsewhere [24]. An
`example of the resulting normalized TEY and AEY
`spectra is shown in Fig. 6. Here we compare spectra of
`an unrubbed BPDA-PDA polyimide, recorded at X-
`ray incidence angles u 908 and u 208. Tests were
`performed regarding radiation damage of the investi-
`gated polymers. At the experimental conditions used
`here, characterized by an X-ray spot size of 0.5 £
`1 mm2 and a sample photo-current of about 5 £
`10211 A at 320 eV, no time dependent changes in
`the spectra were observable.
`
`Fig. 4. LC orientation and pretilt angle in rubbed polyimides and in
`polystyrene. We show only the first monolayer of the LC molecules
`on the rubbed surfaces. More precisely, the LC pretilt angle 1 is
`defined as the average out-of-plane tilt angle of the rods in the bulk
`of the LC.
`
`would allow the development of non-contact alterna-
`tives to the mechanical low-tech rubbing process and
`the development of new, possibly non-polymeric,
`materials which are useful as an alignment layer.
`Below we shall present NEXAFS measurements on
`a variety of rubbed polyimide surfaces and on poly-
`styrene which are shown to provide important new
`insight into the magical alignment process. Polyi-
`mides and polystyrene were chosen for a particular
`reason. As shown in Fig. 4 it is empirically known
`that rubbed surfaces of the two different materials
`align LCs quite differently. All rubbed polyimides
`align LCs along the rubbing direction, with the pretilt
`direction pointing into the rubbing direction. Rubbed
`polystyrene, by contrast, aligns LCs perpendicular to
`the rubbing direction with zero pretilt [8,23]. This
`extreme behavior has to be explained by any reason-
`able model.
`
`3. Experimental details
`
`the
`NEXAFS measurements were performed at
`Stanford Synchrotron Radiation Laboratory on the
`wiggler beam line 10-1 using nearly linearly polarized
`soft-rays with an energy resolution of , 100 meV at
`the carbon K-edge. Spectra were recorded in the
`
`Page 5 of 19
`
`

`
`194
`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`Fig. 5. Parallel (a) and perpendicular (b) experimental geometries used for the NEXAFS measurement. The X-rays are incident on the sample at
`an incidence angle u from the surface. The major component of the electric field vector ~E1 of the elliptically polarized X-rays lies in the
`horizontal plane at the angle u from the sample normal, which is taken as the z axis of the sample frame. The smaller component ~E2 is in the
`vertical direction and lies on the surface of the sample. The sample is rotatable about a vertical axis and around its normal z. The rubbing
`direction is taken along the x axis of the sample frame, as shown in (c). The orientation of the electric field vector and a single molecular p
`orbital in the sample frame are specified by spherical angles, are shown. Because the rubbing process is unidirectional it causes a mirror
`symmetry about the (x,z) plane at the polymer surface, as shown in (d). In general, the molecular symmetry at the surface will only have one-fold
`symmetry about the z axis because the directions x and 2x are inequivalent. As shown in the text one may, however find a molecular frame
`) in which the molecular distribution has at least twofold symmetry about all three axes. This frame is rotated by an angle g about the y (cid:136)
`0
`0
`0
`(x
`,y
`,z
`0
`axis, relative to the sample frame.
`y
`
`We investigated several polyimides and a sample
`of polystyrene. The molecular monomer structures
`are given in Fig. 7. The polyimides were dissolved
`in an organic solvent and spin coated onto 10 £
`10 cm2 ITO coated glass plates to a thickness of
`less than 100 nm. After heating to 858C to remove
`the solvent, the samples were baked at 1808C for
`60 min. Samples were rubbed using a rayon-cloth
`rubbing machine
`at
`200 rpm rotation
`speed,
`25 mm/s plate speed and a pile impression of
`0.6 mm. For the NEXAFS measurements we used
`1 £ 1 cm2 pieces, cut from the unrubbed and rubbed
`sample plates. Polystyrene films of 96 K and 514 K
`molecular weight were dissolved in toluene and spin
`coated onto cleaned Si (100) wafers having a native
`
`oxide layer. Film thicknesses were kept below 100 nm
`to avoid excessive charging in the X-ray beam. The
`samples were heated to 808C to remove the solvent
`and then heated at 1508C for 1 h to relax the films. The
`films were rubbed at room temperature with a velour
`cloth under a load of 2 g/cm2 over a distance of
`300 cm at a speed of 1 cm/s as described elsewhere
`[17,26].
`Before we discuss experimental results we shall
`briefly derive the equation describing the angular
`dependence of the NEXAFS intensity for the case of
`a rubbed polymer surface. Because of
`the low
`symmetry of such systems the angular dependence
`is different
`from that encountered in previous
`NEXAFS studies.
`
`Page 6 of 19
`
`

`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`195
`
`4. Angular dependence of NEXAFS intensity and
`molecular orientation
`
`4.1. NEXAFS intensity in sample frame
`
`As shown in Fig. 6, NEXAFS spectra of polymers
`consist of several peaks or resonances. Because of
`their peak-like shape the lowest-energy resonances
`in the 280–290 eV range are particularly suited for
`an intensity analysis. These resonances arise from
`transitions of a 1s core electron into unoccupied mole-
`cular p* orbitals and their angular dependence
`directly yields the orientation of the molecular p
`system [24]. In principle, the higher-energy s reso-
`nances in the 290–310 eV range could also be used to
`study molecular orientation effects but
`in the
`following we shall restrict our discussion to the p
`resonances. The experimental geometries are speci-
`fied in Fig. 5. For elliptically polarized X-rays the
`angle-dependent NEXAFS intensity I(cid:133)u; F; a; f(cid:134) for
`the p system of a single molecule in the sample frame
`(cid:255)
`(cid:1)
`(x,y,z) is given as [24]
`(cid:255)
`
`I u; F; a; f
`
`(cid:8)
`
`(cid:1)
`
`(cid:1)
`(cid:255)
`(cid:134) I2 u; F; a; f
`
`
`
`(cid:133)
`1 1 2 P
`
`(cid:133)1(cid:134)
`
`:
`
`C P I1 u; F; a; f
`
`Here C is a normalization constant and P is a
`polarization factor describing the relative intensity
`contributions of the orthogonal electric field vector
`components ~E1 and ~E2 of the elliptically polarized
`X-rays [24]. The polarization factor depends on the
`storage ring energy and the beam line optics. Our
`0:80 ^ 0:02,
`experiments were carried out with P
`as discussed below. The general angular depen-
`dence of
`the NEXAFS intensity is derived by
`considering the geometry illustrated in Fig. 5 (c).
`In the figure we show both the electric field vector
`~E of the X-rays and the molecular p orbital as
`vectors. Because the electromagnetic wave oscil-
`lates and the experiment averages over many exci-
`tation events the electric field vector ~E, on average,
`is actually a two-directional vector, so that
`its
`(u,F)
`orientation
`is
`equivalent
`to
`(2u 1 1808; F 1 1808). This
`leads
`to
`the
`following general form of the NEXAFS intensities
`I1 and I2 for the p system of a single molecule in
`
`Fig. 6. Normalized NEXAFS spectra of unrubbed BPDA-PDA
`polyimide are shown for normal (908) and grazing (208) X-ray inci-
`dence angles. Spectra recorded by KLL Auger yield detection are
`shown on top and by means of total electron yield detection at the
`bottom. The spectra are normalized to a unit edge jump in the
`energy range 340–380 eV where all curves coincide. The peaks
`below 290 eV correspond to core electron transitions to unfilled
`p* orbitals, those above 290 eV to s* orbitals in the polymer.
`
`(cid:133)2(cid:134)
`
`Zp
`
`0
`
`Z2p
`(cid:255)
`
`0
`
`(cid:255)
`(cid:1)
`
`(cid:133)4(cid:134)
`In our case the x,z plane is a mirror plane so that
`f a; 2f
`f a; f
`: This
`leads
`to two distinct
`(F 08; 1808)
`and
`‘‘perpendicular’’
`‘‘parallel’’
`(F 908; 2708) experimental geometries as shown
`in Figs. 5 (a) and (b), and the respective NEXAFS
`
`the sample frame (x,y,z), [24]
`I1(cid:133)u; F; a; f(cid:134)
`cos2u cos2a
`1 sin2u sin2a(cid:133)cos2F cos2f1 sin2F sin2f(cid:134)
`sin2u sin2a(cid:133)cosF cosf1 sinF sinf(cid:134);
`1 1
`2
`I2(cid:133)u; F; a; f(cid:134) (cid:136) sin2a(cid:133) sin2F cos2f1 cos2F sin2f(cid:134):
`(cid:133)3(cid:134)
`For a distribution of molecules characterized by a
`distribution function f(a,f) of the p orbitals,
`the
`NEXAFS intensity is given by
`
`I u; F(cid:133) (cid:134)
`(cid:255)
`(cid:1)
`
`(cid:1)
`
`(cid:255)
`
`(cid:1)
`
`f a; f
`
`sina da df:
`
`I u; F; a; f
`
`Page 7 of 19
`
`

`
`196
`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`Fig. 7. Total electron yield NEXAFS spectra in the region of the p resonances for several unrubbed polyimides and for polystyrene, recorded
`for normal (908) and grazing (208) X-ray incidence angles. The changes in fine structure arise from the different chemical environments of the
`carbon atoms in the samples as illustrated by filled circles and discussed in the text. The monomer structures of the polymers are shown on the
`right.
`
`A
`
`sin2u1 C
`
`sin2u;
`
`intensities can be written as,
`(cid:133)5(cid:134)
`k(cid:133)u(cid:134)
`k
`k
`k 1 B
`I
`(cid:133)6(cid:134)
`I’(cid:133)u(cid:134)
`A’ 1 B’sin2u
`where we have defined u to be positive in the (1x,1z)
`and (1y,1z) quadrants and negative in the (2x,1z)
`k
`k
`k
`, A’,
`and (2y,1z) quadrants. The constants A
`, B
`, C
`and B’ depend on the actual distribution function
`k
`and A’ also depend on the polarization
`f a; f
`and A
`factor P.
`
`(cid:255)
`
`(cid:1)
`
`4.2. NEXAFS intensity in molecular frame
`
`In order to specify the average molecular alignment
`in the sample frame and to correlate this alignment
`with that of the LC molecules it is convenient to find
`0
`0
`0
`in which the molecular
`the molecular frame x
`,y
`,z
`distribution is symmetric, i.e. has two-fold or higher
`0
`0
`0
`and z
`axes. For a
`symmetry, with respect to the x
`,y
`
`rubbed surface, the 1x rubbing direction is distinct
`from the 2x direction and the molecular distribution
`has only a onefold rotational symmetry about the
`sample z and x axes, while it has twofold symmetry
`about the y axis. This leads to the different forms of
`Eqs. (5) and (6). In particular, Eq. (5) is seen to be
`asymmetric with respect to u. In realizing that Eq. (5)
`can be written in the form [8],
`(cid:133)7(cid:134)
`k(cid:133)u(cid:134)
`cos2(cid:133)u2 g(cid:134)
`k
`k 1 b
`k 22b
`k
`k
`k 1 b
`cos2g,
`cos2g; B
`where A
`and
`a
`k
`k
`sin2g we see that
`the molecular system
`b
`C
`0
`0
`0
`is simply obtained by rotation of the sample
`,z
`,y
`x
`0
`x,y,z system by an angle g about the y
`axis as
`y
`illustrated in Fig.5(d). The rotation angle is given by,
`k
`(cid:133)8(cid:134)
`
`I
`
`a
`
`k
`
`tan2g
`
`22C
`k
`B
`The angle g is negative for clockwise rotation and
`
`:
`
`Page 8 of 19
`
`

`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`0
`
`197
`
`(cid:133)15(cid:134)
`(cid:133)16(cid:134)
`
`0g:
`
`intensity can always be
`the total
`polarized light
`obtained by measurements along three orthogonal
`directions, independent of the relative orientation of
`the sample coordinate system.
`For the more general case of elliptical polarization
`0
`0
`0
`and a relative rotation of the (x
`) frame by an
`,y
`,z
`0
`angle g about the y
`axis we obtain the following
`y
`equations:
`CfP(cid:133)fx
`0sin2g(cid:134) 1 (cid:133)1 2 P(cid:134)fy
`0g;
`0cos2g1 fz
`0 1 (cid:133)1 2 P(cid:134)(cid:133)fx
`0sin2g(cid:134)g;
`CfPfy
`0cos2g1 fz
`CfP(cid:133)fx
`0cos2g(cid:134)
`0sin2g1 fz
`(cid:133)17(cid:134)
`0sin2g(cid:134)g;
`1 (cid:133)1 2 P(cid:134)(cid:133)fx
`0cos2g1 fz
`(cid:133)18(cid:134)
`CfP(cid:133)fx
`0cos2g(cid:134) 1 (cid:133)1 2 P(cid:134)fy
`0sin2g1 fz
`and I’
`Here I
`z correspond to the measured NEXAFS
`intensities in the parallel and perpendicular geome-
`tries, respectively. These intensities differ because
`the smaller elliptical ~E2 component lies along the y
`and x axes, respectively.
`From these equations the orientation factors are
`derived as:
`
`kz
`
`Ix
`
`Iy
`
`I’
`z
`
`kz
`
`I
`
`!
`
`
`1 1 sin2g
`Pcos2g
`
`k
`
`A’ 1 B
`
`;
`
`(cid:133)19(cid:134)
`
`(cid:133)20(cid:134)
`
`:
`
`(cid:16)
`
`k 1 B’
`
`B
`
`(cid:133)21(cid:134)
`
`0
`
`0 1 fz
`the
`
`1
`total
`
`(cid:133)22(cid:134)
`
`(cid:17)
`
`:
`
`Itot
`
`A
`
`k 1 B’
`Itot
`
`k
`
`A’ 1 B
`
`;
`
`
`
`!
`
`1 2 cos2g
`Pcos2g
`
`(cid:17)
`
`(cid:16)
`
`A
`
`3 2
`
`0
`
`fx
`
`0
`
`fy
`
`0
`
`fz
`
`Itot
`
`Itot
`0 1 fy
`The normalization condition fx
`yields
`the
`following expression for
`integrated intensity Itot
`C
`k 1 A’
`1 3P 2 1
`2P
`The polarization factor can also be directly obtained
`as
`
`k
`B’ 2 B
`k 2 A’ 1 B’ 2 B
`
`P
`
`A
`
`k :
`
`(cid:133)23(cid:134)
`
`y
`
`(cid:1)
`
`positive for anticlockwise rotation about the y
`axis.
`
`(cid:255)
`
`4.3. Molecular orientation factors
`
`(cid:255)
`
`(cid:1)
`
`Since the molecular orientation function f a; f
`cannot be determined by NEXAFS (see Eq. (4)) we
`need a different way to characterize the orientational
`anisotropy. This can be done by simply using three
`orientation factors that describe the relative alignment
`along three orthogonal axes, without actual knowl-
`edge of the orientation function itself [27]. Because
`0
`0
`0
`)
`of the twofold molecular symmetry in the (x
`,y
`,z
`frame the orientation factors are simply the projec-
`tions of f a; f
`along the three axes. This metho-
`dology has previously been extensively used for the
`description of the orientational properties of liquid
`crystals themselves [27,28]. The anisotropy of the
`molecular p system can therefore be described by
`0
`0
`0, fy
`, and fz
`, which are defined as
`orientation factors fx
`the lowest order non-vanishing projections, according
`to
`
`Z
`Z
`Z
`
`0
`
`fx
`
`0
`
`fy
`
`0
`
`fz
`
`sin2a cos2f f(cid:133)a; f(cid:134)dV;
`
`sin2a sin2f f(cid:133)a; f(cid:134)dV;
`(cid:255)
`(cid:1)
`
`cos2a f a; f
`
`dV:
`
`(cid:133)9(cid:134)
`
`(cid:133)10(cid:134)
`
`(cid:133)11(cid:134)
`
`R
`
`(cid:255)
`
`(cid:1)
`
`0
`
`One can envision the orientation factors as the frac-
`0
`0
`0
`, and z
`axes,
`tions of molecules aligned along the x
`, y
`0 1 fy
`0 1 fz
`0
`respectively, and we have fx
`1 for a
`f a; f
`dV:
`normalized distribution
`1) and that
`Assuming linearly polarized light (P
`0
`0
`) are iden-
`the coordinate systems (x,y,z) and (x
`,z
`,y
`tical, we see from Eqs. (2) and (4) that the NEXAFS
`intensities along the x, y and z axes directly determine
`the orientation factors, i.e.
`(cid:133)12(cid:134)
`0 ;
`C fx
`Ix
`(cid:133)13(cid:134)
`(cid:133)14(cid:134)
`Iz
`C fz
`0 1 fz
`0 1 fy
`0
`1 determines
`where the normalization fx
`the constant C to reflect the total integrated NEXAFS
`Ix 1 Iy 1 Iz: Note that for linearly
`intensity C
`Itot
`
`Iy
`
`C fy
`
`0 ;
`
`0 ;
`
`Page 9 of 19
`
`

`
`198
`
`J. Sto¨hr, M.G. Samant / Journal of Electron Spectroscopy and Related Phenomena 98–99 (1999) 189–207
`
`number the C atoms around the phenyl ring in poly-
`styrene beginning with the one attached to the chain,
`we see that atom 1 is different by symmetry from atom
`4 and from atoms 2, 3, 5, and 6, with atoms 3 and 5
`and 2 and 6 being equivalent.
`The spectra of the various polyimides differ signif-
`icantly, owing to the different monomer structures.
`Particularly interesting is the spectrum of BPDA-
`PDA polyimide which contains two different kinds
`of phenyl rings, associated with the central PDA
`group (unshaded) and the BPDA groups (shown
`shaded) in Fig. 7. The BPDA groups form a planar
`system with the CyO bonds so that their p systems are
`parallel and conjugated. Because of conjugation
`effects the CyO p resonance is shifted to lower
`energy by about 0.5 eV relative to that in the other
`two polyimides labelled JSR-1 and NISS-3. Also, the
`phenyl p resonance around 285 eV is composed of
`several overlapping structures. From the

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket