27.06.2013 Views

Linear and nonlinear optical properties of Langmuir–Blodgett ...

Linear and nonlinear optical properties of Langmuir–Blodgett ...

Linear and nonlinear optical properties of Langmuir–Blodgett ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Ž .<br />

Materials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

www.elsevier.comrlocatermsec<br />

<strong>Linear</strong> <strong>and</strong> <strong>nonlinear</strong> <strong>optical</strong> <strong>properties</strong> <strong>of</strong> <strong>Langmuir–Blodgett</strong><br />

multilayers from chromophore-containing maleic acid anhydride<br />

polymers<br />

Sigurd Schrader a,) , Vismants Zauls a , Birgit Dietzel b , Costel Flueraru c , Dietrich Prescher b ,<br />

Jurgen Reiche a , Hubert Motschmann c , Ludwig Brehmer a<br />

¨<br />

a<br />

UniÕersitat ¨ Potsdam, Institut fur ¨ Physik, Lehrstuhl Physik kondensierter Materie, Am Neuen Palais 10, Potsdam D-14469, Germany<br />

b<br />

Institut fur ¨ Dunnschichttechnologie ¨<br />

und Mikrosensorik Teltow, Forschungsgruppe Fluorpolymere, Rudower Chaussee 5, Berlin-Adlersh<strong>of</strong> D-12484,<br />

Germany<br />

c<br />

Max-Planck Institut fur ¨ Kolloid- und Grenzflachenforschung, ¨<br />

Rudower Chaussee 5, Berlin-Adlersh<strong>of</strong> D-12484, Germany<br />

Abstract<br />

Organized polymer films prepared from amphiphilic side-chain polymers were investigated with respect to their application in passive<br />

or active photonic devices. The polymers were synthesized by means <strong>of</strong> polymer analogous reaction <strong>of</strong> azo chromophores with a maleic<br />

acid anhydride polymer which bears saturated hydrophobic side chains. The resulting polymer contains, besides hydrophilic groups <strong>and</strong><br />

hydrophobic side chains, azo chromophores with high second-order hyperpolarizability which are attached to the main chain via a spacer.<br />

In order to increase the hydrophobic character <strong>of</strong> the chromophore, a fluorinated moiety was chosen as acceptor <strong>of</strong> the chromophore.<br />

<strong>Langmuir–Blodgett</strong> Ž LB. technique was applied for preparation <strong>of</strong> multilayer structures with distinct supramolecular architecture from<br />

these polymers. They are characterized by high anisotropy <strong>and</strong> a special combination <strong>of</strong> linear <strong>and</strong> <strong>nonlinear</strong> <strong>optical</strong> Ž NLO. <strong>properties</strong>.<br />

The polymers were deposited both in Y-type <strong>and</strong> Z-type LB layers. The Y-type layers <strong>of</strong> the azopolymer show high birefringence <strong>and</strong><br />

dichroism but small <strong>nonlinear</strong> response which originates from the first deposited layer only. Multilayers <strong>of</strong> the azopolymer deposited in<br />

Y-type configuration are characterized by high internal order which can be concluded from a pronounced Bragg-peak <strong>and</strong> Kiessig-fringe<br />

patterns as obtained from gracing incidence X-ray scattering. Experiments to deposit the active polymer directly in Z-type configuration<br />

by passing a blank water compartment before layer deposition in a double compartment LB-trough led to the result that well-ordered,<br />

anisotropic films are obtained only if the number <strong>of</strong> deposited monolayers is limited to a value <strong>of</strong> about 30. The layers deposited in Z-type<br />

manner show low order. This can be concluded from the fact that only a weak Bragg-peak is observed in the X-ray scattering curves, <strong>and</strong><br />

Kiessig-fringes are very weak or missing. The same conclusion can be drawn from ellipsometric <strong>and</strong> second-harmonic generation Ž SHG.<br />

experiments. A second kind <strong>of</strong> Z-type multilayers was prepared by alternate deposition <strong>of</strong> the active azopolymer <strong>and</strong> <strong>of</strong> another maleic<br />

acid anhydride polymer without chromophores. They show similar linear <strong>optical</strong> <strong>properties</strong> like Y-type layers but due to the<br />

noncentrosymmetric internal structure, a NLO second-order response grows proportional to the square <strong>of</strong> the sample thickness. Again,<br />

thick layers <strong>of</strong> the azopolymer prepared by alternate deposition LB-technique were prepared, which can be used as active waveguide for<br />

second-order NLO application. q 1999 Elsevier Science S.A. All rights reserved.<br />

Keywords: Nonlinear optics; Second-order effects; Fluorine-containing chromophore; Azobenzene; Side-chain polymers<br />

1. Introduction<br />

In order to satisfy the growing dem<strong>and</strong> for a high<br />

b<strong>and</strong>width for <strong>optical</strong> data transfer <strong>and</strong> data processing<br />

systems, a variety <strong>of</strong> passive <strong>and</strong> active components like<br />

polarization-retaining waveguides wx 1 or fast-operating de-<br />

) Corresponding author. Tel.: q49-3319771503; fax: q49-<br />

3319771457; E-mail: sschrader@rz.uni-potsdam.de<br />

0928-4931r99r$ - see front matter q 1999 Elsevier Science S.A. All rights reserved.<br />

Ž .<br />

PII: S0928-4931 99 00088-0<br />

vices like electro-optic amplitude modulators, phase modu-<br />

lators or directional couplers w2–4x is required. Organic<br />

materials are due to their high <strong>optical</strong> <strong>nonlinear</strong>ities <strong>and</strong><br />

their low dielectric constant promising c<strong>and</strong>idates to real-<br />

ize fast <strong>optical</strong> devices w2–6 x.<br />

These devices operate on the<br />

basis <strong>of</strong> second-order <strong>nonlinear</strong> <strong>optical</strong> Ž NLO. effects.<br />

Prerequisites for the appearance <strong>of</strong> a significant macroscopic<br />

second-order NLO susceptibility x 2 in the active<br />

material are a sufficient molecular second-order hyperpolarizability<br />

<strong>and</strong> the presence <strong>of</strong> a noncentrosymmetric in-


528<br />

ternal structure both on the molecular <strong>and</strong> the supramolec-<br />

ular level w2–7 x.<br />

Chromophores with high second-order<br />

hyperpolarizability b are usually linear conjugated chromophores<br />

which bear a strong donor <strong>and</strong> a strong acceptor<br />

in p–p X -position, <strong>and</strong> which have a long, undistorted pelectron<br />

system. Azobenzene chromophores are known to<br />

represent a good compromise between high <strong>nonlinear</strong> response,<br />

<strong>and</strong> other technologically important <strong>properties</strong> like<br />

sufficient solubility in common solvents <strong>and</strong> good photo-<br />

chemical stability w8–18 x.<br />

In order to obtain material <strong>properties</strong><br />

required for waveguide preparation, the chromophores<br />

can be chemically bound to polymeric chains. In<br />

principle, a huge variety <strong>of</strong> materials can be synthesized in<br />

this way which involves main- or side-chain polymers<br />

suitable for spin-coating, amphiphilic polymers for <strong>Langmuir–Blodgett</strong><br />

Ž LB. deposition or polyionenes which can<br />

be used for deposition by self-assembly from polyelec-<br />

trolyte solutions w2–5,19–25x w26,27 x.<br />

In the present study,<br />

amphiphilic side-chain polymers which bear new fluorinecontaining<br />

chromophores were synthesized <strong>and</strong> used for<br />

preparation <strong>of</strong> anisotropic multilyer structures by LB deposition.<br />

Fluorine-containing chromophores exhibit lower refractive<br />

index than fluorine-free chromophores but show,<br />

in case <strong>of</strong> medium or strong donor <strong>and</strong> acceptor sub-<br />

stituents, high b-values w28,29 x.<br />

Therefore, this material<br />

class is <strong>of</strong> special interest for electro-optic <strong>and</strong> other<br />

photonic applications where the figure-<strong>of</strong>-merit is proportional<br />

to the square <strong>of</strong> the second-order <strong>nonlinear</strong> suscepti-<br />

bility <strong>and</strong> to the inverse cube <strong>of</strong> the refractive index w2–4 x.<br />

Many photonic devices are based on the principle <strong>of</strong><br />

<strong>optical</strong> waveguiding w1–4,30–33 x.<br />

A typical waveguide<br />

thickness reaches values around one or a few micrometers.<br />

Preparation <strong>of</strong> waveguides can easily be done by spin-coating<br />

<strong>of</strong> soluble polymers. This technique does not allow for<br />

a fine-tuning <strong>of</strong> internal architecture <strong>of</strong> the waveguide. The<br />

structure <strong>of</strong> waveguides prepared by means <strong>of</strong> LB technique,<br />

however, can be controlled on a molecular scale.<br />

For instance, a gradient <strong>of</strong> refractive index can easily be<br />

introduced through layer-by-layer deposition <strong>of</strong> suitable<br />

amphiphilic polymers. Therefore, the LB deposition technique<br />

is the method <strong>of</strong> choice for fabrication <strong>of</strong> waveguides<br />

having molecularly controlled architecture. This<br />

method involves the compression <strong>of</strong> a monolayer <strong>of</strong> organic<br />

molecules spread on the surface <strong>of</strong> a liquid subphase,<br />

e.g., on water w34 x.<br />

After reaching a sufficient surface<br />

pressure, the monolayer attains a molecular density which<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

resembles the density <strong>of</strong> a two-dimensional solid. By<br />

dipping a pre-treated substrate into the subphase covered<br />

with the compressed monolayer, a material transfer takes<br />

place. During the transfer process, the surface pressure is<br />

kept constant by moving the trough barriers. The transferred<br />

layers have about the thickness <strong>of</strong> the compressed<br />

monomolecular layer on the subphase. Monolayers can be<br />

deposited onto the substrate both on immersion <strong>and</strong> withdrawal,<br />

resulting in a centrosymmetric arrangement <strong>of</strong> both<br />

layers Ž Y-type deposition. or just on immersion ŽX-type<br />

deposition. or on withdrawal Ž Z-type deposition. which<br />

provides noncentrosymmetric structures. The control <strong>of</strong><br />

molecular orientation <strong>and</strong> <strong>of</strong> supramolecular arrangement<br />

is crucial for two aspects — the kind <strong>of</strong> molecular packing<br />

perpendicular to the layer plane is important to obtain<br />

noncentrosymmetric or centrosymmetric structures, while<br />

the molecular orientation with respect to the substrate<br />

plane is important to obtain an anisotropic refractive index.<br />

If the amphiphilic material deposited by LB-technique<br />

is a polymer, then two basic approaches are possible w34 x:<br />

– the polymer is synthesized prior to its deposition <strong>and</strong><br />

then transferred in its compressed state onto the substrate;<br />

or<br />

– the monomeric film is transferred first to the substrate<br />

<strong>and</strong> polymerization takes place after deposition, e.g., by<br />

photochemical polymerization.<br />

For the present study, the first approach was used, i.e.,<br />

the azochromophore-containing polymers were synthesized<br />

<strong>and</strong> characterized before their deposition.<br />

The aim <strong>of</strong> the present study is the evaluation <strong>of</strong> these<br />

polymers with respect to their potential for fabrication <strong>of</strong><br />

passive <strong>and</strong> active waveguide structures by means <strong>of</strong> LB<br />

technique.<br />

2. Materials<br />

2.1. Chromophore synthesis<br />

The functionalized chromophores used for polymer syn-<br />

thesis were 3-w4-Ž 4-trifluoromethyl-phenyl-azo. phenoxyx- propan-1-ol Ž TPPP. <strong>and</strong> 3-w4-Ž4-trifluoromethyl-phenyl- azo. phenoxyx-undecan-1-ol Ž TPPU . . The chromophore,<br />

TPPP, was synthesized according to the scheme given in<br />

Fig. 1, <strong>and</strong> synthesis <strong>of</strong> TPPU was carried out via an<br />

w Ž . x Ž .<br />

Fig. 1. Synthesis <strong>of</strong> the functionalized chromophore, 3- 4- 4-trifluoromethyl-phenylazo -phenoxy -propan-1-ol TPPP .


equivalent route. At first, 4-trifluoromethyl-4 X -hydroxyazobenzene<br />

Ž. 1 was prepared by diazotation <strong>of</strong> trifluoromethyl-aniline<br />

Ž Aldrich. <strong>and</strong> subsequent coupling with<br />

phenol analogues according to a method reported previ-<br />

ously w35 x. Then, equimolar amounts <strong>of</strong> compound Ž. 1 <strong>and</strong><br />

<strong>of</strong> 3-bromopropan-1-ol were heated in dried acetone up to<br />

608C under vigorous stirring for 35 h together with the<br />

fivefold amount <strong>of</strong> dried K 2CO3 <strong>and</strong> a small amount <strong>of</strong><br />

KI.<br />

The inorganic salts were filtered <strong>of</strong>f <strong>and</strong> the volume <strong>of</strong><br />

the filtrate reduced by evaporation in vacuum. By adding<br />

heptane to the organic phase, compound Ž. 2 was precipitated.<br />

The precipitate was filtered <strong>of</strong>f <strong>and</strong> purified by<br />

recrystallization from heptane. An amount <strong>of</strong> 4 mmol<br />

Ž 78% yield. yellow-colored crystals <strong>of</strong> compound Ž 2. were<br />

obtained Ž mp)2408C, decomposed . .<br />

2.2. Polymer synthesis<br />

Polymer synthesis was realized via a polymer analogous<br />

Ž<br />

reaction. One millimole <strong>of</strong> poly maleic anhydrid-co-alt-1-<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537 529<br />

octadecene.Ž. I having a weight average <strong>of</strong> molar mass <strong>of</strong><br />

Mw s 40 000 was converted together with 1 mmol<br />

azochromophore Ž II. Ž3-w4-Ž 4-trifluoromethyl-phenylazo. -<br />

phenoxyx -propan-1-ol. <strong>and</strong> 210 ml Ž three drops. <strong>of</strong> dimethylaminopyridine<br />

as catalyser into a chromophore-containing<br />

side-chain polymer Ž III. Ž cf. Fig. 2 . . The reaction was<br />

carried out in dry acetone under stirring for 30 h at 808C.<br />

In order to precipitate the resulting polymer, the reaction<br />

mixture was subsequently poured onto ice.<br />

The precipitated product was washed several times with<br />

water <strong>and</strong> dried at 508C in vacuum for 4 h. The soluble<br />

by-products were extracted from the polymer with acetone<br />

for 10 h. From the elemental analysis, a fluorine weight<br />

content <strong>of</strong> 3.75% was found for the polymer, which was<br />

insoluble in acetone.<br />

This is equivalent to a conversion rate <strong>of</strong> 30% <strong>of</strong> the<br />

anhydride moieties along the polymer chain. Other analytical<br />

data obtained from infrared Ž IR. investigations support<br />

this result. Therefore, the statistical co-polymer Ž III.<br />

Žazochromophore-containing maleic acid side-chain polymer<br />

— AMS-1. can be described by an average sum<br />

formula as given in Fig. 2 with ms0.3 <strong>and</strong> ns0.7,<br />

Fig. 2. Synthesis <strong>of</strong> the azochromophore-containing side-chain polymer Ž III. Ž AMS-1. from polymer Ž I. <strong>and</strong> chromophore Ž II. Ž3-w4-Ž4-trifluoromethyl- phenylazo. -phenoxyx -propan-1-ol. by polymer analogous reaction.


530<br />

respectively. In the same way, another polymer was synthesized<br />

which differs only in the length <strong>of</strong> the spacer<br />

between chromophore <strong>and</strong> main chain Ž AMS-2 . . This<br />

means that the chromophore, TPPP, is replaced by the<br />

chromophore, TPPU, <strong>and</strong> the polymer, AMS-2, has, consequently,<br />

a spacer <strong>of</strong> 11 CH 2-units<br />

in the side chain.<br />

2.3. Film preparation by LB technique<br />

The LB film deposition was performed on a NIMA 622<br />

alternating trough Ž NIMA Technology Coventry, UK.<br />

equipped with a Wilhelmy plate surface pressure sensor.<br />

The spreading solutions were obtained by dissolving 0.5<br />

mg <strong>of</strong> the polymer in 1.0 ml <strong>of</strong> a N, N-dimethylacetamide<br />

Ž DMAc. rbenzene Ž 1:1. mixture ŽMerck,<br />

high-grade purity<br />

. . The monolayer was formed by spreading 100–200<br />

ml <strong>of</strong> the solution on the water subphase provided by a<br />

Milli-Q system Ž Millipore . . The subphase temperature was<br />

held constant to 218C. The films were compressed with a<br />

constant barrier speed <strong>of</strong> 10 cm2rmin. The surface pressure<br />

— area isotherm <strong>of</strong> polymer Ž III . — is plotted in Fig.<br />

3, where a statistically averaged repeat unit with ms0.3<br />

<strong>and</strong> ns0.7 according to Fig. 2 was assumed for calculation<br />

<strong>of</strong> the area per repeat unit. The monolayer shows a<br />

high collapse pressure <strong>of</strong> about 50 mNrm for both kinds<br />

<strong>of</strong> polymers. They form stable monolayers on water. The<br />

isotherms are reproducible.<br />

For further investigations, LB multilayers have been<br />

prepared on different substrate materials. Multilayer films<br />

<strong>of</strong> the polymers were obtained by depositing the LB<br />

monolayers at a target pressure <strong>of</strong> 25 mNrm <strong>and</strong> a dipper<br />

speed <strong>of</strong> 5 <strong>and</strong> 10 mmrmin. A transfer ratio between 0.8<br />

<strong>and</strong> 1.0 was observed.<br />

Silicon wafers having a native oxide layer <strong>of</strong> 3 nm<br />

thickness, silicon wafers coated with gold by vacuum<br />

evaporation, <strong>and</strong> quartz glass plates used as substrate<br />

Fig. 3. Surface pressure — area isotherm <strong>of</strong> azochromophore-containing<br />

side-chain polymer AMS-1 for the statistically averaged repeat unit <strong>of</strong> the<br />

polymer.<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

material, were hydrophobized by treatment with hexamethyldisilazane<br />

Ž HMDS. prior to deposition. Multilayers <strong>of</strong> 5,<br />

10, 20, 30, 40, <strong>and</strong> 60 monolayers were prepared on these<br />

substrates. These layers were <strong>of</strong> Y-type or Z-type, depen-<br />

dent on the selected way <strong>of</strong> layer deposition w34 x.<br />

In<br />

addition, Z-type layers have been prepared by alternating<br />

deposition <strong>of</strong> two different materials. In that case, alternating<br />

monolayers <strong>of</strong> AMS-1 <strong>and</strong> <strong>of</strong> a chromophore-free<br />

maleic acid anhydride polymer have been deposited in<br />

order to form noncentrosymmetric multilayers.<br />

3. Experimental<br />

The LB multilayers were characterized by X-ray scattering<br />

<strong>and</strong> various linear <strong>and</strong> NLO techniques in order to<br />

obtain information about sample thickness, chromophore<br />

orientation, degree <strong>of</strong> order <strong>and</strong> about their <strong>optical</strong> constants.<br />

3.1. X-ray characterization<br />

X-ray specular reflectiÕity ( XSR) measurements carried<br />

out under gracing incidence provide information about<br />

sample thickness <strong>and</strong> molecular periodicity inside the LB<br />

multilayers. For these measurements, the radiation <strong>of</strong> a<br />

fine focus copper tube with nickel filter Žwavelength<br />

ls<br />

0.154 nm. is used. The radiation is collimated by slit<br />

systems both in the incident <strong>and</strong> the reflected beam. A<br />

goniometer HZG-4 Ž Seifert-FPM, Freiberg, Germany. with<br />

proportional counter was operated in uy2u mode with<br />

u-steps <strong>of</strong> 0.018.<br />

In Fig. 4, the X-ray reflectivity <strong>of</strong> an 18 layer Y-type<br />

LB-film <strong>of</strong> AMS-2 is plotted. A pronounced first-order<br />

001 peak, as well as Kiessig-fringes in the low-angle<br />

region, are visible. The bilayer thickness derived from the<br />

position <strong>of</strong> the 001 peak using the Bragg equation is 3.75<br />

nm. The observation <strong>of</strong> Kiessig-fringes indicates a low<br />

value <strong>of</strong> roughness both <strong>of</strong> the substrate-film <strong>and</strong> the<br />

film–air interface. The overall thickness <strong>of</strong> the film derived<br />

from the Kiessig-fringes is in agreement with a<br />

perfect transfer <strong>of</strong> nine bilayers. Y-type layers <strong>of</strong> AMS-1<br />

are characterized by very similar XSR curves. The bilayer<br />

thickness derived from the position <strong>of</strong> the Bragg-peak is<br />

somewhat higher <strong>and</strong> has a value <strong>of</strong> 3.9 nm. This indicates<br />

a slightly lower density <strong>of</strong> the material which can be the<br />

result <strong>of</strong> a less perfect packing <strong>of</strong> the LB layers. This is<br />

indeed reasonable, since the short-chain spacer <strong>of</strong> AMS-1<br />

causes a stronger coupling to the main chain in comparison<br />

to the polymer AMS-2 where the longer spacer allows a<br />

decoupling between polymer main chain <strong>and</strong> chromophore<br />

packing.<br />

Investigation <strong>of</strong> multilayers <strong>of</strong> AMS-1 deposited in<br />

Z-type configuration shows a different behaviour. The


( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537 531<br />

Fig. 4. X-ray reflectivity curve <strong>of</strong> an 18 layer LB-film <strong>of</strong> the azochromophore-containing polymer, AMS-2, deposited in Y-type configuration on silicon.<br />

X-ray reflectivity curve is characterized by a weak Braggpeak<br />

<strong>and</strong> almost missing Kiessig-fringes. This indicates a<br />

much higher roughness <strong>and</strong> low order in these films. The<br />

reason for this behaviour could be a structural transformation<br />

which occurs during the deposition process.<br />

3.2. Optical characterization<br />

Optical characterization <strong>of</strong> the polymeric LB multilayers<br />

involves both linear <strong>optical</strong> methods like ultraviolet–<br />

visible spectra Ž UV–Vis. absorption spectroscopy <strong>and</strong> ellipsometry,<br />

<strong>and</strong> second-order NLO measurements.<br />

3.3. <strong>Linear</strong> <strong>optical</strong> measurements<br />

<strong>Linear</strong> <strong>optical</strong> measurements on LB samples provide<br />

not only the <strong>optical</strong> material constants but also information<br />

Fig. 5. Ellipsometric angles delta Ž D. <strong>and</strong> psi Ž C . in dependence on the<br />

angle <strong>of</strong> incidence for five monolayers <strong>of</strong> AMS-1 on hydrophobized<br />

silicon Ž squares — experimental values; lines — theoretical curves . .<br />

about the internal structure <strong>of</strong> the material, e.g., the average<br />

thickness <strong>of</strong> a LB monolayer. The two applied techniques<br />

were ellipsometry <strong>and</strong> absorption spectroscopy.<br />

Ellipsometry is a method which is mainly applied for<br />

determination <strong>of</strong> refractive index <strong>and</strong> thickness <strong>of</strong> thin<br />

transparent layers. In this paper, we used a st<strong>and</strong>ard ellipsometer<br />

AFE-400 Ž ZWG, Berlin, Germany. with a rotating<br />

analyser <strong>and</strong> retarder which operates at the wavelength <strong>of</strong><br />

the He–Ne Laser at 632.8 nm. Both azimuth angles <strong>and</strong><br />

four fixed angles <strong>of</strong> incidence between 458 <strong>and</strong> 82.58 are<br />

known with a precision <strong>of</strong> 0.018. Furthermore, free adjustable<br />

angles <strong>of</strong> incidence are known with an accuracy <strong>of</strong><br />

0.058. Two examples <strong>of</strong> spectra <strong>of</strong> multiple angle ellipsometry,<br />

as measured for AMS-1 on different substrates, are<br />

presented in Figs. 5 <strong>and</strong> 6.<br />

Ellipsometric information about refractive index, absorption<br />

<strong>and</strong> thickness <strong>of</strong> the layers under investigation is<br />

Fig. 6. Ellipsometric angles delta Ž D. <strong>and</strong> psi Ž C . in dependence on the<br />

angle <strong>of</strong> incidence for 30 monolayers <strong>of</strong> AMS-1 on gold substrate<br />

Ž squares — experimental values; lines — theoretical curves . .


532<br />

Fig. 7. UV–Vis absorption spectra <strong>of</strong> AMS-1 Z-type multilayers on<br />

quartz glass substrate measured under perpendicular incidence <strong>of</strong> light: 5,<br />

10, 20, 30, 40, 60 monolayers.<br />

condensed into the two ellipsometric angles delta Ž D. <strong>and</strong><br />

psi Ž C . . They express the complex ratio <strong>of</strong> reflectivities <strong>of</strong><br />

S- <strong>and</strong> P-polarized light, i.e., polarized perpendicular <strong>and</strong><br />

parallel to the plane <strong>of</strong> incidence, respectively. Ellipsometric<br />

data were analysed by the use <strong>of</strong> an own simulation <strong>and</strong><br />

data processing program. This program calculates C <strong>and</strong><br />

D values for an assumed input <strong>optical</strong> model <strong>of</strong> the<br />

sample, <strong>and</strong> then optimizes the difference between measured<br />

<strong>and</strong> calculated C <strong>and</strong> D values by varying model<br />

parameters, like layer thicknesses, refractive indices <strong>of</strong> the<br />

different layers or number <strong>of</strong> layers.<br />

The quality <strong>of</strong> fit is judged by the mean square difference<br />

Ž MSD . :<br />

1<br />

MSDs ½ 2 Nypy1<br />

Ž .<br />

=<br />

N<br />

1r2<br />

2 2<br />

Ý Ž c m. i Ž c m.<br />

i 5<br />

i<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

D yD q C yC ,<br />

where D , C <strong>and</strong> D , C are the calculated <strong>and</strong> the<br />

c c m m<br />

measured values <strong>of</strong> the ellipsometric angles, respectively.<br />

N represents the number <strong>of</strong> experimental points <strong>and</strong> p is<br />

the number <strong>of</strong> unknown model parameters. Different points<br />

were measured in order to check the uniformity <strong>of</strong> the<br />

film. In all cases, the fit is accepted if MSD is below 10 y2 .<br />

For measurements carried out at one angle <strong>of</strong> incidence,<br />

the question about the existence <strong>and</strong> the uniqueness <strong>of</strong> the<br />

solution is very important. From the physical point <strong>of</strong><br />

view, a solution obviously exists. The nonuniqueness <strong>of</strong><br />

the solution results from the investigated physical mechanisms<br />

that are potential sources for multiple stationary<br />

solutions. Using different incident angles, it is possible to<br />

find simultaneously the minimum error between simulated<br />

<strong>and</strong> experimental spectra for all angles <strong>of</strong> incidence. Measurements<br />

with different incident angles do not produce<br />

the effect <strong>of</strong> enhancing information content. However, the<br />

multiple-angle measurements will help choose the true<br />

solution.<br />

Absorption spectra <strong>of</strong> LB multilayers on quartz glass<br />

substrates in the UV–Vis range were recorded with a<br />

Perkin-Elmer Lambda 18 spectrophotometer at perpendicular<br />

incidence <strong>of</strong> light. In Fig. 7, the UV–Vis spectra <strong>of</strong><br />

samples having different numbers <strong>of</strong> monolayers are presented.<br />

Three main absorption peaks can be seen as expected<br />

for the azochromophore. The weak absorption b<strong>and</strong> around<br />

450 nm can be assigned to the p–p U transition while the<br />

two strong absorption b<strong>and</strong>s centered at 350 <strong>and</strong> 248 nm<br />

are connected with p–p U transitions <strong>of</strong> the chromophore.<br />

The absorption spectra show an increase <strong>of</strong> absorbance<br />

with the number <strong>of</strong> deposited monolayers.<br />

3.4. NLO measurements<br />

Second-harmonic generation Ž SHG. measurements have<br />

been carried out in order to determine second-order NLO<br />

susceptibility <strong>of</strong> the LB multilayers. Samples were prepared<br />

on quartz glass substrates for transmission measurements<br />

or on silicon or gold substrates for measurements in<br />

reflection geometry. These NLO measurements do not<br />

only provide x 2 values but also information about chromophore<br />

orientation <strong>and</strong> about the internal structure <strong>of</strong> LB<br />

multilayers. The light source for SHG measurements has<br />

been an actively–passively mode-locked Nd:YAG laser<br />

Ž BM Industries, France. which emits laser pulses <strong>of</strong> 30 ps<br />

duration at a wavelength <strong>of</strong> 1064 nm, <strong>of</strong> an energy <strong>of</strong> 7 mJ<br />

per pulse, <strong>and</strong> with a repetition rate <strong>of</strong> 10 Hz. For SHG<br />

measurements in transmission, the sample was rotated on a<br />

computer-controlled rotation stage with the rotation axis<br />

perpendicular to the laser beam.<br />

The detected second-harmonic Ž SH. intensity shows the<br />

typical dependence on the angle <strong>of</strong> incidence ŽMaker<br />

fringes w2–4 x. , as can be seen in Fig. 8 where the SHG<br />

Fig. 8. The SH intensity <strong>of</strong> an LB-film containing alternate monolayers <strong>of</strong><br />

AMS-1 Ž 10 monolayers. <strong>and</strong> octadecyl archidate Ž nine monolayers . . For<br />

comparison, the rescaled signal <strong>of</strong> a quartz reference is also plotted.


Fig. 9. P- <strong>and</strong> S-polarized reflected SHG intensity as function <strong>of</strong> polarization<br />

azimuth <strong>of</strong> fundamental beam. Angle <strong>of</strong> incidence is 458 <strong>and</strong> zero<br />

polarization azimuth corresponds to the S-polarized wave.<br />

intensity <strong>of</strong> a Z-type multilayer <strong>of</strong> AMS-1 deposited in<br />

alternation with a chromophore-free maleic acid anhydride<br />

polymer is plotted vs. the angle <strong>of</strong> incidence. The envelope<br />

<strong>of</strong> this curve contains information about NLO susceptibility<br />

x 2 while the oscillation results from interference <strong>of</strong><br />

harmonic light from layers on both sides <strong>of</strong> the substrate.<br />

The <strong>nonlinear</strong> fit to the envelopes provides the main<br />

tensorial components <strong>of</strong> x 2 <strong>of</strong> the LB multilayer. Details<br />

<strong>of</strong> data analysis are given elsewhere w25 x.<br />

The SHG measurements have been also carried out in<br />

reflection using samples on silicon or gold substrate. For<br />

this kind <strong>of</strong> measurements, the angle <strong>of</strong> incidence <strong>of</strong> the<br />

laser beam was kept constant at 458. The sample could be<br />

rotated with respect to the laser beam, <strong>and</strong> the polarisation<br />

<strong>of</strong> incident light could be varied with respect to the plane<br />

<strong>of</strong> incidence Ž variation <strong>of</strong> polarization azimuth. as well.<br />

The polarization state <strong>of</strong> the incident fundamental beam<br />

was rotated by a Glan polarizer <strong>and</strong> half-wave plate combination.<br />

Reflected SH intensity for P- <strong>and</strong> S-orientations<br />

behind an analyzer has been recorded by a photomultiplier<br />

after spectral selection by an appropriate set <strong>of</strong> filters <strong>and</strong> a<br />

monochromator. A digital storage scope was used for<br />

signal acquisition <strong>and</strong> averaging. Data readout, rotation <strong>of</strong><br />

the sample platform, half-wave plate <strong>and</strong> analyzer have<br />

been performed by a measurement control computer.<br />

A typical result <strong>of</strong> such an experiment is plotted in Fig.<br />

9 where the SH intensity for a sample with five monolayers<br />

<strong>of</strong> AMS-1 on silicon is shown.<br />

The SHG experiments in reflection provide additional<br />

information about the internal structure, especially about<br />

chromophore orientation inside the LB layers. From the<br />

growth <strong>of</strong> SH intensity with the number <strong>of</strong> layers, information<br />

about the type <strong>and</strong> quality <strong>of</strong> chromophore ordering<br />

can be extracted.<br />

A second kind <strong>of</strong> SHG measurements in reflection<br />

keeps the polarization state <strong>of</strong> the exciting laser beam<br />

constant but analyzes the generated harmonic light by the<br />

use <strong>of</strong> an rotating analyzer. This kind <strong>of</strong> experiment pro-<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537 533<br />

vides the same information as the experiment described<br />

before, but it is not reported here.<br />

4. Results <strong>and</strong> discussion<br />

The absorbance <strong>of</strong> the azopolymer multilayers in the<br />

UV–Vis spectral range is plotted in Fig. 7. It depends on<br />

the number <strong>of</strong> monolayers deposited on the substrate in<br />

Z-type. The deposition <strong>of</strong> LB films to solid quartz glass<br />

substrates was carried out with an average transfer ratio <strong>of</strong><br />

0.9 to 1.0. For a transfer ratio <strong>of</strong> 1.0, one would expect an<br />

almost linear increase <strong>of</strong> absorbance with number <strong>of</strong> layers.<br />

This behaviour is, in fact, observed as can be seen in<br />

Fig. 10 where the integrated absorbance <strong>of</strong> the two strong<br />

absorption b<strong>and</strong>s at 350 <strong>and</strong> 248 nm are plotted vs. the<br />

number <strong>of</strong> monolayers.<br />

Since the integrated spectra were not corrected for<br />

reflection losses, the linear plots do not cross the y-axis at<br />

zero but at a value which includes substrate absorption <strong>and</strong><br />

reflection at the different interfaces <strong>of</strong> the sample.<br />

A somewhat different result is obtained from ellipsometric<br />

measurements. Here, an almost linear increase <strong>of</strong><br />

the calculated sample thickness with the number <strong>of</strong> monolayers<br />

is only observed for the first monolayers. Fig. 11<br />

represents the ellipsometrically determined thickness in<br />

dependence on the number <strong>of</strong> AMS-1 monolayers on<br />

silicon substrate. As can be seen, it is not possible to<br />

calculate the thickness <strong>of</strong> monolayers with high accuracy<br />

because the slope <strong>of</strong> this dependence is not constant <strong>and</strong> it<br />

does not cross the origin <strong>of</strong> the graph, too. The same<br />

behaviour is observed for the samples on quartz glass ŽFig.<br />

12. <strong>and</strong> on gold substrates Ž Fig. 13 . . For the latter samples,<br />

the calculated thickness per monolayer amounts to 1.6 nm<br />

if the range between 5 <strong>and</strong> 10 monolayers is considered.<br />

For samples on silicon Ž Fig. 11. <strong>and</strong> quartz glass substrates<br />

Ž Fig. 12 . , the calculated thicknesses per monolayer<br />

are 1.7 <strong>and</strong> 1.3 nm, respectively.<br />

Fig. 10. Integrated absorbance <strong>of</strong> the two strong absorption b<strong>and</strong>s <strong>of</strong><br />

AMS-1 at 350 <strong>and</strong> 248 nm in dependence on the number <strong>of</strong> monolayers.


534<br />

Fig. 11. Thickness <strong>of</strong> AMS-1 multilayers on silicon in dependence on the<br />

number <strong>of</strong> monolayers determined by ellipsometry at a wavelength <strong>of</strong><br />

632.8 nm.<br />

The value ns1.557 <strong>of</strong> the refractive index was found<br />

to be the same for these layers, independent <strong>of</strong> the kind <strong>of</strong><br />

substrate.<br />

For samples with more than 20 monolayers, the thickness<br />

per monolayer, as calculated above, cannot be used<br />

since no reasonable solution can be found for thicker<br />

samples on the base <strong>of</strong> these parameters. As a consequence,<br />

a new ellipsometric model with two virtual layers<br />

has been used. The first virtual layer is built up <strong>of</strong> 10<br />

monolayers <strong>of</strong> polymer AMS-1, <strong>and</strong> the next layer covers<br />

the remaining number <strong>of</strong> monolayers deposited on the<br />

substrate. This procedure was applied to the various samples<br />

on different substrates. It was found that the thickness<br />

per monolayer for the upper virtual layer is about three<br />

times smaller than that for the lower virtual layer. Our<br />

evaluation for samples on three different substrates led to<br />

the conclusion that, the thickness per monolayer decreases<br />

with increasing number <strong>of</strong> monolayers. If the number <strong>of</strong><br />

monolayers exceeds 40, the film starts to collapse —<br />

Fig. 12. Thickness <strong>of</strong> AMS-1 multilayers on quartz glass in dependence<br />

on the number <strong>of</strong> monolayers.<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

Fig. 13. Thickness <strong>of</strong> AMS-1 multilayers on gold in dependence on the<br />

number <strong>of</strong> monolayers.<br />

which means that the next monolayers have a thickness<br />

below 0.2 nm in terms <strong>of</strong> the applied model which includes<br />

a third virtual layer covering all monolayers above<br />

40. In Table 1, the results are summarised.<br />

In most cases, the LB deposition is sensitive to the<br />

temperature dependence <strong>of</strong> surface pressure <strong>of</strong> the floating<br />

monolayer, to the composition <strong>of</strong> the subphase, <strong>and</strong> to the<br />

surface <strong>properties</strong> <strong>of</strong> the substrate. In order to reduce the<br />

number <strong>of</strong> unknown parameters, the LB deposition was<br />

always carried out under the same conditions. Therefore,<br />

the change in the deposition behaviour with increasing<br />

number <strong>of</strong> monolayers should have its origin in material-<br />

inherent <strong>properties</strong>. In fact, Robinson et al. w36x observed a<br />

similar behaviour for 22-tricoseneic acid <strong>and</strong> arachidic<br />

acid. Using contact angle measurements, he confirmed the<br />

change in surface free energy relative to the substrate for<br />

successive layers. The surface free energy for 22-tricoseneic<br />

acid <strong>and</strong> arachidic acid decreases with increasing<br />

deposition pressure until triple-line pinning occurs. Then<br />

an increase <strong>of</strong> surface energy with number <strong>of</strong> monolayers<br />

has been measured. Hence, with increasing number <strong>of</strong><br />

monolayers, the influence <strong>of</strong> substrate decreases, which<br />

leads to a loss <strong>of</strong> molecular order especially in the upper<br />

layers. It is, therefore, not possible to deposit oriented<br />

multilayers from this kind <strong>of</strong> organic molecule alone,<br />

because the deposition is, in this case, sensitive to the<br />

surface energy which depends on thickness.<br />

Table 1<br />

Thickness per monolayer for Z-type multilayers <strong>of</strong> AMS-1 dependent on<br />

the number <strong>of</strong> monolayers deposited on different kinds <strong>of</strong> substrates<br />

determined by multiple angle ellipsometry<br />

Substrate Sample with different numbers <strong>of</strong> monolayers Ž nm.<br />

5 10 20 30 40 60<br />

Silicon 1.64"0.12 0.55"0.04 about 0.1<br />

Gold 1.70"0.25 0.55"0.13 about 0.1<br />

Quartz 1.29"0.24 0.52"0.01 about 0.1


Fig. 14. Reflected SHG <strong>of</strong> Z-type AMS-1 multilayers on silicon in<br />

dependence on the number <strong>of</strong> monolayers. The expected N 2 -dependence<br />

is shown by the dashed line.<br />

In case <strong>of</strong> AMS-1 Z-type multilayers, the chromophores<br />

<strong>of</strong> the first 10 monolayers are oriented almost perpendicular<br />

to the substrate plane, but they become more <strong>and</strong> more<br />

tilted as the number <strong>of</strong> monolayers increases. Obviously, a<br />

similar situation, as described by Robinson, is valid also<br />

for the present case. The reason could be that the polar<br />

groups <strong>of</strong> the polymer main chain cannot compensate the<br />

large dipole moments <strong>of</strong> the chromophores when the interaction<br />

with the substrate loses its influence. As a result,<br />

disorder in the upper layers increases in comparison to the<br />

lower layers.<br />

The investigation <strong>of</strong> molecular order by SHG led to the<br />

same result. The reflected SHG intensity <strong>of</strong> an adsorbate–<br />

surface system can schematically be given as a sum <strong>of</strong><br />

contributions from substrate surface <strong>nonlinear</strong> susceptibil-<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537 535<br />

ity x 2 s <strong>and</strong> from the susceptibility <strong>of</strong> deposited multilayers<br />

x 2 m:<br />

< 2 2 2 < 2<br />

I2 v A xs qxm I v .<br />

For perfectly aligned LB multilayer, x 2 m is proportional<br />

to the number <strong>of</strong> deposited layers N <strong>and</strong> as a result, a<br />

dependence according to I AN 2 2 v should be observable<br />

( 2 v<br />

as linear increase in a I plot vs. N. SHG reflection<br />

measurement were carried out by the use <strong>of</strong> a circularly<br />

polarized fundamental light beam <strong>and</strong> without an analyzer<br />

in the detector path in order to average contributions <strong>of</strong> all<br />

x 2 tensor components. The increase <strong>of</strong> the SHG signal<br />

m<br />

follows the expected N 2-dependence only for the first 30<br />

monolayers, as shown by the ( I2 v vs. N plot in Fig. 14.<br />

Comparing this dependence with the results <strong>of</strong> ellipsometric<br />

measurements, it is clear that orientation <strong>and</strong> ordering<br />

<strong>of</strong> chromophores are increasingly reduced for Z-type<br />

multilayers above a thickness <strong>of</strong> 30 monolayers.<br />

This is consistent with results <strong>of</strong> SHG experiments<br />

dependent on the polarization <strong>of</strong> the incident laser beam.<br />

For a sample <strong>of</strong> five monolayers Ž Z-type, AMS-1 . , the<br />

detected SH signal in S- <strong>and</strong> P-polarisation in dependence<br />

on the polarization angle <strong>of</strong> incident fundamental beam is<br />

plotted as polar diagram in Fig. 15. The observed intensity<br />

is described by the following expression w37 x:<br />

ž /<br />

2p 2 sec 2 2 v<br />

u<br />

2 l´ 0 c<br />

< ˆ2<br />

v Ž .<br />

2<br />

s ˆv Ž . ˆv<br />

Ž<br />

2 < 2 . v<br />

I s e L 2v<br />

Px :e L v e L v I ,<br />

where l is the wavelength, c is the speed <strong>of</strong> light, ´<br />

st<strong>and</strong>s for vacuum permittivity, the term, e ˆV , is a unit<br />

polarization vector <strong>and</strong> LŽ V . are appropriate linear local<br />

field correction factors at frequencies Vsv <strong>and</strong> Vs2v,<br />

respectively.<br />

Ž .<br />

Fig. 15. Reflected SH intensity <strong>of</strong> a Z-type AMS-1 multilayer sample five monolayers in dependence on the polarizer azimuth <strong>of</strong> the incident<br />

fundamental beam. Both S- <strong>and</strong> P-polarized SH intensities are plotted.


536<br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537<br />

Fig. 16. Reflected SH intensity <strong>of</strong> a Z-type AMS-1 multilayer covering 60 monolayers.<br />

Model calculations show that this intensity pattern resembles<br />

a situation where the chromophore is tilted towards<br />

the normal <strong>of</strong> the substrate plane by an angle <strong>of</strong><br />

about 308. The intensity distribution for a sample with 60<br />

monolayers is shown in Fig. 16.<br />

The features, as visible in Fig. 15, are still present as<br />

fingerprint <strong>of</strong> the lower layers but this signal is superimposed<br />

by a second signal <strong>of</strong> layers having a clearly different<br />

average tilt angle, which deviates considerably from<br />

the normal direction to the substrate plane, indicating a<br />

more disordered state. A detailed analysis will be pub-<br />

lished elsewhere w38 x.<br />

Since multilayers <strong>of</strong> AMS-1 prepared as Z-type layers<br />

collapse if a large number <strong>of</strong> layers is deposited, the<br />

second approach is used where alternating layers <strong>of</strong> the<br />

azopolymer <strong>and</strong> <strong>of</strong> the chromophore-free maleic acid anhydride<br />

polymer are prepared. Effective NLO coefficients<br />

Ž 2 d s0.5x . <strong>of</strong> these multilayer films have been deter-<br />

33 33<br />

mined by means <strong>of</strong> Maker-fringe technique. A typical<br />

pattern, as received from SHG measurements, is shown in<br />

Fig. 9. Films were deposited on quartz glass substrates <strong>and</strong><br />

SHG measurements were made in transmission for P–P<br />

<strong>and</strong> S–P polarization geometry. The signal from 1-mm<br />

thick quartz reference plate was used for calibration. We<br />

found d values <strong>of</strong> about 10 pmrV for our samples<br />

33<br />

having 30 or less monolayers. Investigations on thicker<br />

multilayer samples are still in progress.<br />

5. Conclusions<br />

The LB multilayers, prepared from side-chain polymers<br />

<strong>of</strong> maleic acid anhydride which carry fluorine-containing<br />

NLO azochromophores in the side chain, show high second-order<br />

NLO susceptibility if they are deposited as<br />

Z-type layers. These layers can be prepared from the<br />

azopolymer alone or by alternating deposition with other<br />

materials, e.g., with a chromophore-free maleic acid anhydride<br />

polymer. Multilayers prepared from the azopolymer<br />

alone tend to collapse if the number <strong>of</strong> layers exceeds a<br />

value <strong>of</strong> 30, which is confirmed both by linear <strong>and</strong> by<br />

NLO measurements. The ellipsometrically determined<br />

thickness per monolayer reduces gradually for increasing<br />

number <strong>of</strong> deposited layers, <strong>and</strong> the second-order NLO<br />

response grows less rapid with sample thickness than<br />

expected. Both results reveal that the ‘‘Z-type’’ layers <strong>of</strong><br />

the pure azopolymer are characterized by high disorder<br />

which can be due to structural rearrangement during the<br />

process <strong>of</strong> layer deposition. This behaviour can be explained<br />

by insufficient screening <strong>of</strong> the strong chromophore<br />

dipoles, <strong>and</strong> related to that by strong changes <strong>of</strong><br />

surface energy with increasing thickness. Alternating LB<br />

layers prepared from the azopolymers <strong>and</strong> other amphiphiles<br />

are promising c<strong>and</strong>idates for fabrication <strong>of</strong> thicker<br />

LB layers with high anisotropy <strong>and</strong> high NLO susceptibility.<br />

Investigations on these materials are still in progress.<br />

Typical values <strong>of</strong> <strong>nonlinear</strong> bulk <strong>optical</strong> coefficients d 33<br />

are in the range <strong>of</strong> 10 pmrV for LB multilayers prepared<br />

from azochromophore-containing maleic acid anhydride<br />

polymer <strong>and</strong> a respective chromophore-free polymer by<br />

alternating Z-type deposition.<br />

Acknowledgements<br />

Financial support <strong>of</strong> the Deutsche Forschungsgemeinschaft<br />

under project number Schr 462r3-1, <strong>and</strong>r3-2, <strong>and</strong><br />

project Kn 439r2-1 <strong>and</strong> <strong>of</strong> Ministry <strong>of</strong> Education, Culture<br />

<strong>and</strong> Science <strong>of</strong> the Country <strong>of</strong> Br<strong>and</strong>enburg is gratefully<br />

acknowledged.


References<br />

wx 1 G.P. Agrawal, Nonlinear Fiber Optics, 2nd edn., Academic Press,<br />

San Diego, 1995, 592 pp.<br />

wx 2 P.N. Prasad, D.J. Williams, Introduction to Nonlinear Optical Effects<br />

in Molecules <strong>and</strong> Polymers, Wiley, New York, 1991.<br />

wx 3 C.-C. Tang, High-speed electro-optic modulators from <strong>nonlinear</strong><br />

<strong>optical</strong> polymers, in: H.S. Nalwa, S. Miyata Ž Eds. . , Nonlinear Optics<br />

<strong>of</strong> Organic Molecules <strong>and</strong> Polymers, CRC Press, Boca Raton, FL,<br />

1997, pp. 441–464.<br />

wx 4 M.G. Kuzyk, C.W. Dirk Ž Eds. . , Characterization Techniques <strong>and</strong><br />

Tabulations for Organic Nonlinear Optical Materials, Marcel Dekker,<br />

New York, 1998.<br />

wx 5 D.J. Williams, Organic polymeric <strong>and</strong> nonpolymeric materials with<br />

large <strong>optical</strong> <strong>nonlinear</strong>ities, Angew. Chem., Int. Ed. Engl. 23 Ž 1984.<br />

690.<br />

wx 6 D.S. Chemla, J. Zyss Ž Eds. . , Nonlinear Optical Properties <strong>of</strong> Organic<br />

Molecules <strong>and</strong> Crystals, Vols. 1 <strong>and</strong> 2, Academic Press, New York,<br />

1987.<br />

wx 7 R. Boyd, Nonlinear Optics, Academic Press, San Diego, 1992.<br />

wx 8 M. Barzoukas, A. Fort, G. Klein, C. Serbutoviez, L. Oswald, J.F.<br />

Nico, Conformational dependence <strong>of</strong> quadratic hyperpolarizabilities<br />

<strong>of</strong> a series <strong>of</strong> push–pull stilbenes — characterization <strong>and</strong> investigation<br />

<strong>of</strong> empirical correlations, Chem. Phys. 64 Ž.Ž 3 1992. 395–406.<br />

wx 9 L.-T. Cheng, W. Tam, S.H. Stevenson, G.R. Meredith, G. Rikken,<br />

S.R. Marder, Experimental investigations <strong>of</strong> organic molecular <strong>nonlinear</strong><br />

<strong>optical</strong> polarizabilities: 1. Methods <strong>and</strong> results on benzene <strong>and</strong><br />

stilbene derivatives, J. Phys. Chem. 95 Ž 26. Ž 1991. 10631–10643.<br />

w10x L.-T. Cheng, W. Tam, S.R. Marder, A.E. Stiegman, G. Rikken,<br />

C.W. Spangler, Experimental investigation <strong>of</strong> organic molecular<br />

<strong>nonlinear</strong> <strong>optical</strong> polarizabilities: 2. A study <strong>of</strong> conjugation dependence,<br />

J. Phys. Chem. 95 Ž 26. Ž 1991. 10643–10652.<br />

w11x K. Clays, A. Persoons, Hyper-Rayleigh scattering in solution, Phys.<br />

Rev. Lett. 66 Ž 23. Ž 1991. 2980.<br />

w12x K. Clays, A. Persoons, Hyper-Rayleigh scattering in solution, Rev.<br />

Sci. Instrum. 63 Ž.Ž 6 1992. 3285–3289.<br />

w13x C.R. Moylan, S.A. Swanson, C.A. Walsh, J.I. Thackava, J.R. Twieg,<br />

R.D. Miller, V.Y. Lee, From EFISH to electro-optic measurements<br />

<strong>of</strong> <strong>nonlinear</strong> chromophores, Proc. SPIE 2025 Ž 1993. 192–201.<br />

w14x C.R. Moylan, R.J. Twieg, V.Y. Lee, S.A. Swanson, K.M. Betterton,<br />

R.D. Miller, Nonlinear <strong>optical</strong> chromophores with large hyperpolarizabilities<br />

<strong>and</strong> enhanced thermal stabilities, J. Am. Chem. Soc. 115<br />

Ž 26. Ž 1993. 12599–12600.<br />

w15x J.L. Oudar, Optical <strong>nonlinear</strong>ities <strong>of</strong> conjugated molecules: stilbene<br />

derivatives <strong>and</strong> highly polar aromatic compounds, J. Chem. Phys. 67<br />

Ž.Ž 2 1997. 446–457.<br />

w16x J.L. Oudar, H. LePerson, Second-order polarizabilities <strong>of</strong> some<br />

aromatic molecules, Opt. Commun. 15 Ž 29. Ž 1975. 258–262.<br />

w17x M.A. Pauley, H.W. Guan, C.H. Wang, A.K.Y. Jen, Determination <strong>of</strong><br />

first hyperpolarizability <strong>of</strong> <strong>nonlinear</strong> <strong>optical</strong> chromophores, J. Chem.<br />

Phys. 104 Ž 20. Ž 1996. 7821.<br />

w18x P.C. Ray, P.K. Das, Dissociation constants <strong>of</strong> weak organic acids in<br />

protic solvents obtained from the first hyperpolarizabilities in solution,<br />

J. Phys. Chem. 99 Ž 51. Ž 1995. 17891.<br />

w19x M. Amano, T. Kaino, Second-order <strong>nonlinear</strong>ity <strong>of</strong> novel diazo-dye<br />

attached polymer, J. Appl. Phys. 68 Ž 12. Ž 1990. 6024.<br />

w20x L.R. Dalton, A.W. Harper, R. Ghosn, W.H. Steier, M. Ziavi, H.<br />

Fellerman, Y. Shi, R.V. Mustacich, A.K.Y. Jen, K.J. Shen, Synthesis<br />

<strong>and</strong> processing <strong>of</strong> improved organic second-order <strong>nonlinear</strong><br />

( )<br />

S. Schrader et al.rMaterials Science <strong>and</strong> Engineering C 8–9 1999 527–537 537<br />

<strong>optical</strong> materials for applications in photonics, Chem. Matter 7<br />

Ž 1995. 1060–1081.<br />

w21x A.K.Y. Jen, V.P. Rao, K.J. Drost, Y.M. Cai, R.M. Mininni, J.T.<br />

Kenney, E.S. Binkley, L.R. Dalton, S.R. Marder, Progress on heteroaromatic<br />

chromophores in high temperature polymers for<br />

electro-optic applications, Proc. SPIE 2143 Ž 1994. 30–40.<br />

w22x C.R. Moylan, R.J. Twieg, V.Y. Lee, R.D. Miller, W. Volksen, J.L.<br />

Thachava, C.A. Walsh, Synthesis <strong>and</strong> characterization <strong>of</strong> thermally<br />

robust electro-optic polymers, Proc. SPIE 2285 Ž 1994. 17–30.<br />

w23x K.D. Singer, H.E. Katz, J.E. Sohn, C.W. Dirk, L.A. King, H.M.<br />

Gordon, Second-order <strong>nonlinear</strong> <strong>optical</strong> <strong>properties</strong> <strong>of</strong> donor- <strong>and</strong><br />

acceptor-substituted aromatic compounds, J. Opt. Soc. Am. B 6<br />

Ž 1989. 1339.<br />

w24x S. Schrader, K. Pfeiffer, F. Reuther, H.-J. Lorkowski, Stable polymer<br />

chromophore systems for <strong>nonlinear</strong> optics, in: E. Silinsh, A.<br />

Medvids, A. Lusis, A. Ozols Ž Eds. . , Proc. SPIE — Optical Organic<br />

<strong>and</strong> Semiconducting Inorganic Materials, Vol. 2968, 1997, pp. 61–<br />

66.<br />

w25x S. Schrader, D. Prescher, V. Zauls, New chromophores <strong>and</strong> polymers<br />

for second-order <strong>nonlinear</strong> optics, in: M. Eich Ž Ed. . , Second-<br />

Order Organic Nonlinear Optics, Proc. SPIE, Vol. 3474, 1998, pp.<br />

160–171.<br />

w26x M. Florsheimer, ¨ M. Kupfer, ¨ C. Bosshard, H. Looser, P. Gunter, ¨<br />

Phase-matched <strong>optical</strong> second-harmonic generation in Langmuir–<br />

Blodgett film waveguides by mode conversion, Adv. Mater. 4 Ž 12.<br />

Ž 1992. 795–798.<br />

w27x G. Decher, F. Klinkhammer, Makromol. Chem. Macromol. Symp.<br />

46 Ž 1991. 19.<br />

w28x S. Schrader, D. Prescher, Molecular substitution <strong>and</strong> <strong>nonlinear</strong> <strong>optical</strong><br />

<strong>properties</strong> <strong>of</strong> chromophores, in: E. Silinsh, A. Medvids, A. Lusis,<br />

A. Ozols Ž Eds. . , Proc. SPIE — Optical Organic <strong>and</strong> Semiconductor<br />

Inorganic Materials, Vol. 2968, 1997, pp. 44–49.<br />

w29x S. Schrader, D. Prescher, K. Lukaszuk, R. Wortmann, A.H. Otto,<br />

Solvatochromy <strong>and</strong> electro-<strong>optical</strong> study <strong>of</strong> new fluorine-containing<br />

chromophores, in: M. Eich Ž Ed. . , Second-Order Organic Nonlinear<br />

Optics, Proc. SPIE, Vol. 3474, 1998, pp. 14–22.<br />

w30x C.N. Ironside, J.S. Aitchison, J.M. Arnold, An all-<strong>optical</strong> switch<br />

employing the cascaded second-order <strong>nonlinear</strong> effect, IEEE J.<br />

Quantum Electron. 29 Ž 10. Ž 1993. 2650–2654.<br />

w31x Y. Baek, R. Schiek, G.I. Stegeman, G. Krijnen, I. Baumann, W.<br />

Sohler, All-<strong>optical</strong> integrated Mach–Zehnder switching due to cascaded<br />

<strong>nonlinear</strong>ities, Appl. Phys. Lett. 68 Ž 15. Ž 1996. 2055–2057.<br />

w32x R. Schiek, Y. Baek, G.I. Stegeman, Second-harmonic generation <strong>and</strong><br />

cascaded <strong>nonlinear</strong>ity in titanium-undiffused lithium niobate channel<br />

waveguides, J. Opt. Soc. Am. B 15 Ž.Ž 8 1998. 2255–2268.<br />

w33x L. Palchetti, S. Sottini, D. Gr<strong>and</strong>o, E. Giogetti, R. Ricceri, G.<br />

Gabriellli, Electro-optic modulation <strong>of</strong> laser diode by mode interference<br />

in a multilayer waveguide including a 2-docosylamino-5nitropyridine<br />

<strong>Langmuir–Blodgett</strong> film, Appl. Phys. Lett. 72 Ž. 8<br />

Ž 1998. 873.<br />

w34x R.H. Tredgold, Order in Thin Organic Films, Cambridge Univ.<br />

Press, Cambridge, 1994.<br />

w35x D. Prescher, T. Thiele, R. Ruhmann, G. Schulz, J. Fluorine Chem.<br />

74 Ž 1995. 185.<br />

w36x I. Robinson, J.-R. Sambles, N.A. Cade, Thin Solid Films 178 Ž 1989.<br />

125.<br />

w37x Y.R. Shen, The Principles <strong>of</strong> Non-linear Optics, Wiley, New York,<br />

1992.<br />

w38x V. Zauls, S. Schrader, C. Flueraru, H. Motschmann, in preparation.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!